51
Research Division Federal Reserve Bank of St. Louis Working Paper Series The Economic and Statistical Value of Forecast Combinations under Regime Switching: An Application to Predictable US Returns Massimo Guidolin and Carrie Fangzhou Na Working Paper 2006-059B http://research.stlouisfed.org/wp/2006/2006-059.pdf November 2006 Revised April 2007 FEDERAL RESERVE BANK OF ST. LOUIS Research Division P.O. Box 442 St. Louis, MO 63166 ______________________________________________________________________________________ The views expressed are those of the individual authors and do not necessarily reflect official positions of the Federal Reserve Bank of St. Louis, the Federal Reserve System, or the Board of Governors. Federal Reserve Bank of St. Louis Working Papers are preliminary materials circulated to stimulate discussion and critical comment. References in publications to Federal Reserve Bank of St. Louis Working Papers (other than an acknowledgment that the writer has had access to unpublished material) should be cleared with the author or authors.

The Economic and Statistical Value of Forecast ... · 2For a special case, Aiolfi and Timmermann (2005) provide conditions under which the population MSFE of an equally-weighted

  • Upload
    others

  • View
    2

  • Download
    0

Embed Size (px)

Citation preview

Page 1: The Economic and Statistical Value of Forecast ... · 2For a special case, Aiolfi and Timmermann (2005) provide conditions under which the population MSFE of an equally-weighted

Research Division Federal Reserve Bank of St. Louis Working Paper Series

The Economic and Statistical Value of Forecast Combinations under Regime Switching:

An Application to Predictable US Returns

Massimo Guidolin and

Carrie Fangzhou Na

Working Paper 2006-059B http://research.stlouisfed.org/wp/2006/2006-059.pdf

November 2006 Revised April 2007

FEDERAL RESERVE BANK OF ST. LOUIS Research Division

P.O. Box 442 St. Louis, MO 63166

______________________________________________________________________________________

The views expressed are those of the individual authors and do not necessarily reflect official positions of the Federal Reserve Bank of St. Louis, the Federal Reserve System, or the Board of Governors.

Federal Reserve Bank of St. Louis Working Papers are preliminary materials circulated to stimulate discussion and critical comment. References in publications to Federal Reserve Bank of St. Louis Working Papers (other than an acknowledgment that the writer has had access to unpublished material) should be cleared with the author or authors.

Page 2: The Economic and Statistical Value of Forecast ... · 2For a special case, Aiolfi and Timmermann (2005) provide conditions under which the population MSFE of an equally-weighted

The Economic and Statistical Value of ForecastCombinations under Regime Switching: AnApplication to Predictable US Returns∗

Massimo Guidolin†

University of Manchester Business School and

Federal Reserve Bank of St. Louis

Carrie Fangzhou Na

Fannie Mae

Abstract

We address one interesting case — the predictability of excess US asset returns from macroeconomic

factors within a flexible regime switching VAR framework — in which the presence of regimes may lead

to superior forecasting performance from forecast combinations. After having documented that forecast

combinations provide gains in prediction accuracy and these gains are statistically significant, we show

that combinations may substantially improve portfolio selection. We find that the best-performing fore-

cast combinations are those that either avoid estimating the pooling weights or that minimize the need for

estimation. In practice, we report that the best-performing combination schemes are based on the prin-

ciple of relative, past forecasting performance. The economic gains from combining forecasts in portfolio

management applications appear to be large, stable over time, and robust to the introduction of realistic

transaction costs.

JEL codes: C53, E44, G12, C32.

Keywords: Forecast Combination, Predictability, Multivariate Regime Switching, Portfolio Performance.

1. Introduction

One of the least debated stylized facts in the modern empirical finance literature is that the laws of motion

governing the returns on most assets are subject to recurrent “breaks”, discrete shifts in the dynamic structure

generating risk premia, volatility, correlations, and — at least in principle — all relevant properties of financial

returns. For instance, Paye and Timmermann (2006) find four major breaks in the past fifty years in a large

set of size- and industry-sorted US equity portfolios, as well as 18 international stock market portfolios, with

breaks occurring in different time periods.

Although other, competing approaches exist, one of the most widespread ways in which such evidence is

translated into precise modeling strategies consists of adopting models in the regime switching class.1 Following

∗Elizabeth La Jeunesse provided excellent research assistance.We are indebted to thank Mark Wohar (a co-editor) and one

anonymous referee for a number of stimulating comments and suggestions. We also thank session partecipants at the St. Louis

University conference on “Forecasting in the Presence of Structural Breaks and Model Uncertainty” (August 2006).†Correspondence to: Prof. Massimo Guidolin, Manchester Business School, Accounting & Finance Division. Phone: +44-

(0)161-306-6406; Fax: +44-(0)161-275-4023. Address: University of Manchester Business School, MBS Crawford House, Booth

Street East, Manchester M13 9PL, United Kingdom. E-mail: [email protected] and Krolzig (1999) discuss regime switching models as ways to capture structural change.

Page 3: The Economic and Statistical Value of Forecast ... · 2For a special case, Aiolfi and Timmermann (2005) provide conditions under which the population MSFE of an equally-weighted

Hamilton (1989), several papers have proposed to improve the empirical fit of standard, single-equation models

for short-term interest rates (e.g., Gray (1996) and Ang and Bekaert (2002b)) and stock returns (e.g., Turner,

Starz, and Nelson (1989) and Ang and Bekaert (2002a)) by allowing for mixtures of distributions. For instance,

Turner, Starz, and Nelson (1989) develop a univariate model with regime shifts in means and variances,

showing that mean excess equity returns tend to be low in the high-risk (volatility) period, and viceversa.

Allowing for switching in the parameters of an autoregressive conditional heteroskedasticity (ARCH) process,

Hamilton and Susmel (1994) report that in-sample performance and out-of-sample forecasts of the regime-

switching ARCH are superior to a benchmark single-state GARCH(1,1) specification and that the high-

volatility state is likely to occur in recession periods. Guidolin and Timmermann (2006a) extend this class of

models to multivariate systems including excess returns on a few equity portfolios as well as bond returns.

Up to this point, the empirical finance literature has developed following an intuitive research agenda: to

uncover evidence of regimes in financial data, to propose models that provide an accurate in-sample fit, and

to assess their usefulness for prediction and financial decision making, e.g. portfolio choice, asset pricing, and

risk management. For instance, Guidolin and Ono (2006) estimate a range of multivariate regime switching

VAR models for a rich eight-variable vector that includes stock and bond returns in excess of a T-bill rate,

the T-bill yield, typical predictors used in finance (such as the default spread and the dividend yield), and

three macroeconomic variables: inflation, industrial production growth, and real money growth. After finding

evidence of four regimes and of time-varying covariances, they show that the recursive, out-of-sample predictive

performance of the four-state model is superior to a simpler (and nested) VAR(1).

However, the forecasting literature has made it clear that in the presence of structural instability in

econometric relationships it is highly unlikely that a single, non-linear model — by necessity, only a misspecified

simplification of the true but unknown data generating process — may provide the most accurate forecasts. On

the contrary, forecasting experts have for a long time been aware that forecast combinations can be considered

as a hedge against non-stationarities. For instance, Winkler (1989) argues that

“(...) in many situations there is no such thing as a ‘true’ model for forecasting purposes. The

world around us is continually changing, with new uncertainties replacing old ones.”

Elliott and Timmermann (2004) argue that different individual forecast models may react quite differently

to structural breaks, such as institutional or policy changes. Some models, such as the ones allowing for regime

shifts, are only temporarily affected by the breaks and adapt more rapidly than other models. A combination

that puts more weight on adaptive models before or after the breaks and loads more heavily on stable models

in the middle is expected to give better results than either of the two alone. This means that the existence of

structural shifts of unknown form/timing makes it likely not that some non-linear framework able to capture

such features would manage to produce the best out-of-sample performance, but that on the contrary pooling

forecasts may be the winning strategy.2

Our paper has two objectives. First, to present an interesting case study — the predictability of excess US

asset returns from macroeconomic factors within the flexible regime switching VAR framework investigated by

Guidolin and Ono (2006) — in which the presence of regimes may lead to superior forecasting performance from

forecast combinations. Second — after having shown that in practice forecast combinations do provide gains

2For a special case, Aiolfi and Timmermann (2005) provide conditions under which the population MSFE of an equally-

weighted combined forecast will be lower than the population MSFE of the best model. Clements and Hendry (1998) illustrate

the usefulness of the forecast combination approach for artificial data subject to a structural break.

2

Page 4: The Economic and Statistical Value of Forecast ... · 2For a special case, Aiolfi and Timmermann (2005) provide conditions under which the population MSFE of an equally-weighted

in prediction accuracy and that these gains are statistically significant — to investigate the economic value of

such gains. We do that by assessing the improvement in (risk-adjusted) portfolio performance made available

to a portfolio manager who — confronted with the evidence of regimes and the availability of a multiplicity of

forecasting models — entertains the use of a number of forecast combination schemes.

Our main results can be summarized as follows. First, we extend and confirm Guidolin and Ono’s (2006)

findings that carefully modeling regime switches in the econometric relationships tying excess asset returns to

macroeconomic and financial factors provides both a better dynamic description of the joint density of the data

as well as a superior set of prediction tools.3 Second, we show that forecast combinations may substantially

improve the prediction accuracy relative to a large set of heterogeneous models. Such improvements turn out to

be strong enough to carry high levels of statistical significance. Third, we find that an important qualification

must be applied to our claim: the best performing forecast combinations are those that either avoid estimating

the pooling weights (i.e. simple, equally weighted forecasts) or — better — those that minimize the need for

estimation and, therefore, the estimation error “absorbed” by the forecaster. In practice, we report that

especially at short horizons, the best performing combination schemes are based on the principle of relative,

past forecasting performance (see e.g., Stock and Watson, 2003) and as such are poised to avoid estimating

covariances between past forecasting performances across different models. This confirms the classical adage

in forecasting that the major concern one should have about pooling forecasts involves the way to estimate the

weights assigned to each individual model. A potential drawback of forecast combinations associated with this

is the estimation error induced by the weight estimation. Fourth, we report large and statistically significant

economic gains from combining forecasts for portfolio management purposes. In particular, we show that

— depending on the risk aversion coefficient — such a value may be identified with a 8.4% increase in mean

annualized portfolio returns or an increase in the realized Sharpe ratio from 0.20 to 0.36. Both measures seem

hardly negligible and may represent powerful incentives for money managers to seriously entertain pooling

the predictions yielded by a number of competing models as a serious alternative to simply picking one “best

performing” candidate over the others. Finally, such gains turn out to be stable over time — i.e. not to

depend on any speficic sub-sample of the overall 1985-2004 period — and robust to taking transaction costs

into account.

Our paper contributes to two distinct literatures. On one hand, there are now numerous studies that have

found that forecast combinations — even simple mean (equally weighted), trimmed mean, or median schemes —

tend to outperform the best individual forecasts.4 The general principle is that almost all econometric speci-

fications only provide a local approximation to the real data generating process. The extreme complication in

the return series makes it impossible for a single model to be superior to others all the time. If the information

incorporated in two (sets of) models are not completely overlapping, the combination of the two may give

better forecasts than either alone. These ideas are particularly relevant in the presence of non-stationarities,

which are notoriously difficult to effectively model without recurring to approximations.

On the other hand, an impressive literature has piled up that investigates whether US stock and bond

returns are predictable using past values of macroeconomic variables. In fact, using regression models, numer-

3Notice that in a forecasting perspective there appears to be no clear consensus as to whether allowing for non-linearities may

concretely lead to an improved forecasting performance (see e.g. De Gooijer and Kumar, 1992, Clements and Krolzig, 1998).

Stock and Watson (2001) report that non-linear models tend to outperform linear ones only at short horizons. The gains are

however small.4General discussion of forecast combinations and additional references may be found in Clemen (1989), Stock and Watson

(2001), Diebold and Lopez (1996), Hendry and Clements (2002), and Timmermann (2005).

3

Page 5: The Economic and Statistical Value of Forecast ... · 2For a special case, Aiolfi and Timmermann (2005) provide conditions under which the population MSFE of an equally-weighted

ous studies have found that in each data set a few macroeconomic variables can be found that systematically

predict US stock returns. Among others, Fama and French (1988) document that the dividend yield forecasts

future returns on common stocks. Fama and Schwert (1977) report that real stock returns are negatively

related with expected and unexpected components of inflation, and that industrial production and real GNP

growth have forecasting power for financial returns. Campbell (1987) presents evidence that a variety of term

structure variables such as two- and six-month spreads as well as the 1-month T-bill rate, all forecast excess

stock and bond returns. Fama and French (1989) confirm this result using data at alternative frequencies

and a longer sample period (1927-1987). Similarly, Fama and French (1989) show that the default spread is a

significant predictor of stock returns. Fama and French (1993) extend this evidence to excess bond returns.

A related paper is Guidolin and Timmermann (2005b), in which optimal combination weights are derived

within a four-state regime switching model that captures common latent factors driving short-term US spot

and forward rates. In their framework, the combination weights explicitly depend on the state of the economy.

They find that combining forecasts from different models helps improve the out-of-sample forecasting perfor-

mance for short-term interest rates in the US. Apart from obvious differences in the application, our paper

differs from theirs because it does not impose the presence of regimes when computing weights, but simply

uses realized forecast errors to uncover the optimizing weights. Moreover, while Guidolin and Timmermann

(2005b) refrain from an assessment of the economic value of their regime-switching forecast combinations, in

our paper we attach a precise price tag to the usefulness of combinations in portfolio management.

The paper is structured as follows. Section 2 describes the forecasting models. Section 3 gives information

on the data employed in the paper. Section 4 estimates a range of regime switching models and proceeds to

select the one providing the best fit according to a number of statistical criteria. Parameter estimates and

interpretation for the resulting regimes are provided. Section 5 shows that a four-state model produces useful

out-of-sample forecasts. Section 6 introduces the forecast combination problem, comments on the recursive

combination weights derived applying standard methods, and gives results on the relative performance of

combinations in out-of-sample experiments. Section 7 evaluates the economic value of forecast combinations

by focussing on a real time, pseudo out-of-sample portfolio choice experiment. Section 8 concludes.

2. The Models

Suppose that the random vector collecting monthly (excess) returns on l different assets (the vector xt) and

q macroeconomic variables (mt), possibly predicting (and predicted by) asset returns, follows a k−regimeMarkov switching (MS) V AR(p) process with heteroskedastic components, compactly MSIAH(k, p) (see

Krolzig, 1997):5

yt = μSt +

pXj=1

Aj,Styt−j +ΣSt²t (1)

with ²t ∼ NID(0, Il+q) and yt ≡ (x0t m0t)0.6 St is a latent state variable driving all the parameters in equation

(1). μSt collects the l + q regime-dependent intercepts, while the (l + q) × (l + q) matrix ΣSt represents

the factor in a state-dependent Choleski factorization of the covariance matrix, ΩSt . A non-diagonal ΣSt

captures simultaneous co-movements. Moreover, dynamic (lagged) linkages across different asset markets and

5I, A and H refer to state dependence in the intercept, vector autoregressive terms and covariance matrix. p is the autoregressive

order. Models in the class MSIH(k, 0)-V AR(p) have regime switching in the intercept but not in the VAR coefficients.6Assume the absence of roots outside the unit circle, thus making the process covariance stationary. Ang and Bekaert (2002b)

show that for covariance stationarity to obtain, it is sufficient for such a condition to be verified in at least one of the k regimes.

4

Page 6: The Economic and Statistical Value of Forecast ... · 2For a special case, Aiolfi and Timmermann (2005) provide conditions under which the population MSFE of an equally-weighted

between financial markets and macroeconomic influences are captured by the VAR(p). In fact, conditionally

on the unobservable state St, (1) defines a standard Gaussian, reduced-form VAR(p) model. On the other

hand, when k > 1, alternative hidden states are possible that will influence both the conditional mean and

the volatility/correlation structure characterizing (1). The states are generated by a discrete, homogeneous,

irreducible and ergodic first-order Markov chain:7

Pr³St = j|Sjt−1j=1, yjt−1j=1

´= Pr (St = j|St−1 = i) = pij , (2)

where pij is the generic [i, j] element of the k × k transition matrix P. Ergodicity implies the existence of a

stationary vector of probabilities ξ satisfying ξ = P0ξ.

(1) nests a number of simpler model in which either some parameter matrices are not needed or become

independent of the regime. The simpler models may greatly reduce the number of parameters to be estimated.

Among them, we will devote special attention to MSIH(k, 0)-V AR(p) models,

yt = μSt +

pXj=1

Ajyt−j +ΣSt²t, (3)

a special case of equation (1) in which intercepts and covariance matrix are regime-dependent, while the

VAR(p) coefficients are not. This restricted sub-class of models turns out to be important in the following.

A limit case of (1) is obtained when k = 1, a standard multivariate Gaussian VAR(p) model:

yt = μ+

pXj=1

Ajyt−j +Σ²t. (4)

Three additional models are entertained in the following: Pesaran and Timmermann (1995)-style predictive

regressions, Box-Jenkins style univariate ARIMA models, and simple “no-change” benchmarks (see e.g., Neely

and Weller, 2000) for excess asset returns. Pesaran and Timmermann’s (PT) regressions imply that at time t

xt = μx +

pxXj=1

Bjyt−j +Σxut ut ∼ NID(0, Il), (5)

i.e. past values of both excess returns and of macroeconomic variables predict subsequent excess returns.

Taken at face value, equation (5) simply corresponds to the first l rows of equation (4). In fact, Pesaran

and Timmermann (1995) experiment with a number of statistical criteria to recursively select which of the

variables in yt ought to be included in the predictive regression. In what follows we experiment with two such

criteria, the selection of variables from yt which either (i) maximize the adjusted R2, or (ii) minimize the BIC

criterion.

As for the univariate ARIMA class, we restrict ourselves to consider Gaussian AR(p) models, which are

special cases of (5),

xnt = φn0 +

pnXj=1

φnj xnt−j + σnunt unt ∼ NID(0, 1), (6)

7The assumption of a first-order Markov process is not restrictive, since a higher order Markov chain can always be repa-

rameterized as a higher dimensional first-order Markov chain. On the other hand, relaxing the assumption of fixed transition

probabilities (see, e.g., Diebold, Rudebusch, and Sichel, 1993) complicates estimation but could lead to improved forecasting

performance. We leave this extension for future research.

5

Page 7: The Economic and Statistical Value of Forecast ... · 2For a special case, Aiolfi and Timmermann (2005) provide conditions under which the population MSFE of an equally-weighted

(n = 1, ..., l) in which restrictions are imposed on the coefficient matrices Bj . The number of lags is selected

by minimizing the BIC criterion.8

The no change benchmarks correspond to (5) when Bj = O for j = 1, ..., px, i.e. when xt = μx +Σxut

which implies the following (approximate) Gaussian random walk with drift for log-asset prices:

ln(Pt) = ln(Pt−1) + rf ιl + μx +Σxut = d+ ln(Pt−1) +Σxut (7)

which is derived from the definition xt ≡ ln(Pt)− ln(Pt−1)− rf ιl where rf is the risk-free interest rate and

ιl is a l × 1 column vector of ones. In this case, μx is the risk premium, assumed to be constant over time.

Notice that in principle, the criteria of variable selection within PT’s framework may endogenously yield

(7), although in general we would expect this not to be the case. equation (7) has played a key role in the

development of the modern theory of asset pricing and portfolio choice, while it also corresponds to a natural

random walk yardstick popular in forecasting and applied time series analysis. Therefore it seems natural to

compare our models — and forecast combinations built on them — to this constant risk premium benchmark.9

3. The Data

We use monthly data for the period 1926:12 - 2004:12, a total of 937 observations per variable. Asset returns

are from the Center for Research on Security Prices (CRSP) at the University of Chicago. In particular, we

use data on the three most important segments of the US market: stocks (value-weighted stock returns for the

NYSE, NASDAQ, and the AMEX exchanges), bonds (a CRSP index of 10-year to maturity US government

bonds), and money market instruments (30-day Treasury bills, again from CRSP).10

Additionally, we employ five predictor variables which either correspond to macroeconomic aggregates or

that have been associated to business cycle conditions in earlier research. In the first group we have the CPI

inflation rate (seasonally adjusted), the rate of growth of industrial production (seasonally adjusted), and the

rate of growth of a measure of adjusted monetary base. These three series are available from FREDR°II at

the Federal Reserve Bank of St. Louis. The practice of seasonally adjusting the data in real time experiments

(See sections 5-7) assumes that market participants are effectively able to ‘see through’ the veil of time series

variation purely caused by seasonal factors. In the latter group we have two variables. The first is the dividend

yield (calculated from CRSP data), defined as aggregate dividends on the value-weighted CRSP portfolio of

stocks over the previous twelve month period divided by the current stock price. The second is the default

spread, defined as the differential yield on Moody’s Bbb (low rating) and Aaa (high rating) seasoned corporate

bonds with similar maturities. These two variables have played a key role in the recent literature on optimal

asset allocation under predictable asset returns, see e.g. Fugazza, Guidolin, and Nicodano (2006).

In our empirical analysis we study excess stock returns, defined as the difference between nominal, realized

monthly returns and the 1-month T-bill rate. For similar reasons, given the important literature on term

spreads in the US yield curve, we use the long-short bond term spread (the term premium) defined as the

excess return of CRSP long-term bond over 1-month T-bill returns. Finally, also the monetary base growth

is measured in real terms, by deducting from nominal rates of growth the realized CPI inflation rate.

8Notice that settling on some value pn ≥ 2 does not impose that φnj 6= 0 for j < pn. For instance a model such as xnt =

φn0 + φn1xnt−1 + φn3x

nt−3 + σnunt might be selected.

9Coulson and Robins (1993) advocate using “no-change” (random walk) models in forecast combination experiments.10The bond returns data are completed by using the Ibbotson-Sinquifeld data for the period 1926-1946.

6

Page 8: The Economic and Statistical Value of Forecast ... · 2For a special case, Aiolfi and Timmermann (2005) provide conditions under which the population MSFE of an equally-weighted

Tables 1 reports basic summary statistics. Mean values are consistent with commonly known facts: for

instance, the mean excess stock return is 0.65% per month, i.e. 7.80% per year, which represents a typical

value in the equity premium literature, with an annualized volatility of 19.1%; the mean term premium is

0.14% per month, i.e. 1.68% per year, a moderate but plausible average slope of the US term structure;

the average annualized real T-bill rate is 0.60% which, summed to a mean annualized inflation rate of 3.12%,

delivers a mean annualized nominal short-term interest rate of 3.72%, once more in line with the typical values

reported in the asset pricing literature. Both real money and industrial production growth are positive on

average, 0.48 and 2.52 percent in annualized terms, respectively.

Table 1 approximately here

All the series display evident departures from a (marginal) Gaussian distribution, which would imply

zero skewness (i.e. a symmetric distribution) and a kurtosis coefficient of 3. On the opposite, both excess

stock returns and all macroeconomic variables are characterized by huge kurtosis values (in excess of 10), an

indication of distributions with tails considerably fatter than a normal. The dividend yield has only moderate

kurtosis, but it is also skewed to the right (which is to be expected, since the dividend yield cannot be

negative by construction). Even in the case of excess bond returns, a formal Jarque-Bera test of marginal

normal distribution rejects with a 0.000 p-value. For all series (the exception is excess bond returns, which are

not serially correlated) there is evidence of statistically significant first-order serial correlation, as evidenced

by Ljung-Box statistics (of order 4) in excess of the 1% critical value under a χ2(4). Similarly, there is evidence

of volatility clustering (heteroskedasticity), as all Ljung-Box statistics (of order 4) applied to squared values

of the variables are highly significant.

4. Econometric Estimates

This section reviews some of the results in Guidolin and Ono (2006) and contains the bulk of our estimation,

in-sample results. Section 4.1 presents a number of model selection criteria and shows that a four-state model

in which the VAR component of the model is time-homogeneous outperforms a number of competing models

in many dimensions. Section 4.2 presents parameter estimates for such a model. The sub-section establishes

that a regime switching model is required to provide an accurate in-sample fit to the data.

4.1. Model Selection

We estimate a large number of variants of (1) and use five alternative criteria to gauge of the correct spec-

ification of the candidate models. In particular, we estimate models with k = 1, 2, 3, 4, with and without

regime-dependent covariance matrix, and with VAR coefficients that may or may not depend on the regime.11

The first selection criterion concerns the appropriate number of regimes k in model (1). In particular,

we would like to test whether the null of a single-state model (k = 1) can be rejected in favor of k > 1.

As already stressed, when k = 1 equation (1) reduces to a simpler Gaussian VAR(p) model. As discussed

in Garcia (1998), testing for the number of states in a regime switching framework may be tricky. Given

some k ≥ 2, the problem is that under any number of regimes smaller than k there are a few parameters of

11We refrain from trying and estimating models with k ≥ 5 since the number of parameters quickly grows to levels that makeeither estimation uncertainty overwhelming or that cause the MLE-EM routines to fail. For instance, a MSIAH(5,1) model implies

560 paramters, i.e. only 13.4 observations per parameter.

7

Page 9: The Economic and Statistical Value of Forecast ... · 2For a special case, Aiolfi and Timmermann (2005) provide conditions under which the population MSFE of an equally-weighted

the unrestricted model — e.g., some (or all) elements of the transition probability matrix associated to the

rows that correspond to “disappearing states” – that can take any values without influencing the likelihood

function. We say that these parameters become a nuisance to the estimation. The result is that the presence

of nuisance parameters gives the likelihood surface so many degrees of freedom that computationally one can

never reject the null that the non-zero values of those parameters were purely due to sampling variation. This

implies that asymptotically the LR statistic fails to have a standard chi-square distribution. Davies (1977)

circumvents the nuisance parameters problem by deriving an upper bound for the significance level of the LR

test under nuisance parameters:

Pr (LR > x) ≤ Pr¡χ21 > x

¢+√2x exp

³−x2

´ ∙Γ

µ1

2

¶¸−1.

where Γ (·) is the standard gamma function. We find that even adjusting for the presence of nuisance parame-ters, the evidence against specifying traditional single-state models is overwhelming: the smallest LR statistic

takes a value of 139, which is clearly above any conceivable critical value regardless of number of restrictions

imposed.

Once we establish that k ≥ 2 is appropriate, this only rules out models of simpler VAR type. We thereforeproceed to select an appropriate model within the more general regime switching class MSIAH(k, p) with

k ≥ 2. As in a few other applied papers on regime switching models (e.g., Sola and Driffill (1994) and Guidolinand Timmermann (2006c)), we employ a battery of information criteria, i.e. the Akaike (AIC), Bayes-Schwartz

(BIC), and Hannan-Quinn (H-Q) criteria. These criteria are supposed to trade-off in-sample fit with prediction

accuracy. In practice, information criteria identify the ex-ante potential for good out-of-sample performance

by penalizing models with a large number of parameters. We find evidence of some tensions among different

criteria. The AIC is minimized by a richly parameterized MSIAH(4,1) model in which 444 parameters have

to be estimated. Although MLE estimation could be carried out, issues may exist with a model that implies

a saturation ratio (i.e. the number of available observations per estimated parameter) of only 16.9. However,

this is less than surprising as the AIC is generally known to select large models in nonlinear frameworks (see

e.g., Fenton and Gallant (1996)). Next, the H-Q seems to be undecided between a relatively parsimonious

MSIH(4,0)-VAR(1) model (with saturation ratio of almost 30) and a richer MSIAH(3,1) (saturation ratio of

23). Notice that these two models imply a different number of regimes, 3 vs. 4. So, if on one hand it seems

obvious that regime switching matters, the precise number of states required remains debatable. Finally, the

BIC selects a relatively tight MSIH(4,0)-VAR(1) model.

All in all, we are left with two plausible and competing candidate models. The first one is a four-regime

MSIH(4,0)-VAR(1) model that is directly selected by both the H-Q and the parsimonious BIC criterion. The

second is a three-regime MSIAH(3,1) model that obtains a good ‘score’ in a H-Q metric.12 As discussed by

Guidolin and Ono (2006), these two models are structurally different both in a statistical and in an economic

sense. In particular, MSIAH(3,1) implies regime switching in the VAR(1) coefficients, i.e. in this model the

dynamic linkages between financial markets and the macroeconomy are time-varying; this is not the case for

the MSIH(4,0)-VAR(1) model.

Finally, density specification tests are used to “break the impasse” between MSIAH(3,1) and MSIH(4,0)-

VAR(1). Regime switching models consists of flexible mixtures which − if the number of regimes k is expanded12We do not pursue estimation of the richer MSIAH(4,1) (selected by the AIC) because of the high probability of it being

over-parameterized (its saturation ratio is half the MSIH(4,0)-VAR(1)).

8

Page 10: The Economic and Statistical Value of Forecast ... · 2For a special case, Aiolfi and Timmermann (2005) provide conditions under which the population MSFE of an equally-weighted

with the sample size − may be thought of as providing a seminonparametric approximation of the processfollowed by the joint conditional density of the data, see Marron and Wand (1992). In this framework, it

has become customary to require of a regime switching model that they provide a correct specification for

the entire conditional distribution of the variables at hand. The seminal work of Diebold et al. (1998) has

spurred increasing interest in specification tests based on the h-step ahead accuracy of fit of a model for

the underlying density. These tests are based on the probability integral transform or z-score. This is the

probability of observing a value smaller than or equal to the realization of yt+1, yt+1, under the null that the

model is correctly specified. Under a k-regime mixture of normals, this is given by (n = 1, ..., l +m)

Pr¡ynt+1 ≤ ynt+1|=t

¢=

kXi=1

Φ

⎛⎝σ−1n,i

⎡⎣ynt+1 − μn,i − e0npX

j=1

Aj,iyt+1−j

⎤⎦⎞⎠Pr(St+1 = i|=t) ≡ znt+1,

where σn,i is the volatility of variable n in state i, and en is a vector with a one in position n and zeros

elsewhere. If the model is correctly specified, znt+1 should be independently and identically distributed (IID)

and uniform on the interval [0, 1].13

Unfortunately, testing whether a distribution is uniform is not a simple task, as tests popular in the

statistics literature often rely on the IID-ness of the series, which is here at stake as well. Therefore Berkowitz

(2001) has recently proposed a likelihood-ratio test that inverts Φ to get a transformed z-score (dropping the

n superscript),

z∗t+1 ≡ Φ−1(zt+1),

which essentially transforms the z-score back into a bell-shaped random variable. Provided that the model

is correctly specified, z∗ should be IID and normally distributed (NID(0, 1)). To test this hypothesis, we

use a likelihood ratio test that focuses on a few salient moments of the return distribution. Suppose the

log-likelihood function is evaluated under the null that z∗t+1˜ NID(0, 1):

LIIN(0,1) ≡ −T

2ln(2π)−

TXt=1

(z∗t )2

2.

Under the alternative of a misspecified model, the log-likelihood function incorporates deviations from the

null, z∗t+1˜ NID(0, 1):

z∗t+1 = α+

qXj=1

lXi=1

βji(z∗t+1−i)

j + σut+1, (8)

where ut+1 ∼ NID(0, 1). The null of a correct return model implies q × l + 2 restrictions — i.e., α = βji = 0

(j = 1, ..., q and i = 1, ..., l) and σ = 1 — in equation (8). Let L(α, βjiq lj=1 i=1, σ) be the maximized log-

likelihood obtained from equation (8). To test that the null model (some version of (1)) is correctly specified,

we can then use the following test statistic:

LRql+2 ≡ −2hLIIN(0,1) − L(α, βji

q lj=1 i=1, σ)

id→ χ2ql+2.

In addition to the standard Jarque-Bera test that considers skew and kurtosis in the z-scores to detect non

normalities in z∗t+1, it is customary to present three likelihood ratio tests, namely a test of zero-mean and

13The uniform requirement relates to the fact that deviations between realized values and projected (fitted) ones should be

conditionally normal and as such describe a uniform distribution once it ‘filtered through’ an appropriate Gaussian cdf. The IID

condition reflects the fact that if the model is correctly specified, errors ought to be unpredictable and fail to show any detectable

structure.

9

Page 11: The Economic and Statistical Value of Forecast ... · 2For a special case, Aiolfi and Timmermann (2005) provide conditions under which the population MSFE of an equally-weighted

unit variance (q = l = 0), a test of lack of serial correlation in the z-scores (q = 1 and l = 1) and a test

that further restricts their squared values to be serially uncorrelated in order to test for omitted volatility

dynamics (q = 2 and l = 2). Notice that a rejection of the null of normal transformed z-scores has the same

meaning as rejecting the null of a uniform distribution for the raw z-scores, i.e. the model fails in generating

a density with the appropriate shape. A rejection of the zero-mean, unit variance restriction points to specific

problems in the location and scale of the density underlying the model. A rejection of the restriction that

z∗t+1 is IID (i.e. the presence of serial correlation in levels or squares) points to dynamic misspecifications.Strikingly, a simple and yet popular VAR(1) is resoundingly rejected by all tests and for all variables.

Rejections tend to be harsh: the highest VAR(1) p-value is 0.001, i.e. there is actually a very thin chance that

the data might have been generated by a simple linear Gaussian homoskedastic model. In fact, the rejections

are so strong that it becomes difficult to understand in which direction one should be moving to amend the

VAR(1) model to improve the in-sample performance.14

The picture improves, albeit not drastically, when a MSIAH(3,1) model is estimated. For most tests and

variables, the LR test statistics decline by a factor between 30 and 200% when we move from a single- to a

multi-state model. The exceptions are few (the Jarque-Bera tests for excess bond returns, the inflation rate,

and adjusted monetary base growth). However. all (but one, the real T-bill rate and when testing the zero-

mean unit-variance properties) of the related p-values remain highly significant, indicating strong rejection of

the null of correct specification of the three-state model.

Figure 1 provides visual evidence on the sources of misspecifications within a MSIAH(3,1) model by

displaying quantile-quantile (q-q) plots for the empirical distribution of z∗t+1 for each of the eight variables.Notice that if z∗t+1 is N(0, 1), the q-q should approximately look like a 45 degrees straight line in the q-qplane. This seems to happen only for real T-bill yields. On the other hand, a few plots assume an S-shape,

i.e. the slope is too high at the center of the distribution (i.e. more mass is put under the distribution of

z∗t+1 than under a N(0, 1)) and too flat for intermediate values in the support (where mass is missing vs.

the Gaussian case). This is the case of excess stock and bond returns, inflation, IP and monetary growth

rates. In two other cases − dividend yield and the default spread − the q-q plots are flatter than a 45 degreesline, a sign that too much mass is simply moved to the tails of the corresponding distributions.

Figure 1 approximately here

On the contrary, the improvement is strong and significant when we fit a four-state model. The p-values

associated to the various tests generally increase and out of 32 combinations test/variable, we have that the

null of no misspecification fails to be rejected in 15 cases, with p-values exceeding 0.05. Of the remaining 17

tests, in 7 the p-values are between 0.01 and 0.05, i.e. the rejection is rather mild. For 3 variables − remarkablyall of the financial variables, including the dividend yield − the tests give evidence of correct specification,with only some concerns caused by the potential of additional volatility clustering not accommodated by

regime switching covariance matrices in excess stock returns, and by deviations of the shape of the bond and

dividend yield scores from normality. The improvement vs. the three-state model is clear: in only one case

the LR statistic increases when the number of regimes is increased (the test for correct location and scale

of the distribution of the transformed z-scores) and the variation in the corresponding p-value is moderate.

Figure 2 presents almost perfect q-q plots, i.e. roughly aligned around a 45 degrees line.

14Detailed results are reported in Guidolin and Ono (2006).

10

Page 12: The Economic and Statistical Value of Forecast ... · 2For a special case, Aiolfi and Timmermann (2005) provide conditions under which the population MSFE of an equally-weighted

Figure 2 approximately here

4.2. A Four-State Model

Table 2 presents parameter estimates. Panel A reports estimates of a benchmark, single-state VAR(1) model.

Panel B shows MLE-EM estimates of the four-state model. Panel A shows that in a standard VAR(1) many of

the estimated coefficients are not significant. The implications for predictability of financial returns are rather

interesting: apart from a weak own serial correlation (coefficient is 0.10), excess stock returns are essentially

unpredictable using any of the macro instruments entertained in the paper. The same is true for excess bond

returns, which can be just (weakly) predicted from past excess stock returns (coefficient -0.02). Much more

predictability characterizes real short-term interest rates, which are (as expected) highly persistent (coefficient

0.58) and can also be predicted off past default spreads (coefficient 1.28) and IP growth (-0.06). However,

only limited confidence should be attributed to these results for three reasons: we know from Section 4.1 that

single-state VAR(1) models are rejected even when account is taken of nuisance parameter issues. If there are

multiple regimes in the data, we can expect that all estimates obtained from single-regime models might be

severely biased.15

Table 2 approximately here

Things greatly improve under a well-specified regime switching model. Table 2, panel B, starts by showing

that the fraction of parameters in the conditional mean function that are precisely estimated grows when

multiple states are allowed. For instance, most of the intercepts parameters are now highly significant. However

the most visible changes concern indeed the amount (and in some cases, the structure) of the predictability

patterns implied by the model. On one hand, modeling regimes erases all traces of own- and cross-serial

correlation involving excess asset returns. This is not surprising as structural breaks (regimes) are well known

to artificially inflate the degree of own-persistence. On the other hand, excess stock and bond returns become

now highly predictable using lagged values of three variables: real T-bill rates (which forecast lower excess

returns, since the real short-term rate enters the discount rate in asset pricing models) as in Campbell (1987),

the default spread (which forecasts higher excess returns, as in Fama and French (1989)), and the inflation

rate (which forecasts lower future excess returns, presumably as a consequence of the recessions that need to

be induced to bring inflation under control) as in Fama and Schwert (1977). Excess stock returns are also

predicted by past IP growth (as in Cutler, Poterba and Summers (1989)), although the economic effect is

small.16

Table 2 also reports the regime-dependent estimated volatilities and pairwise correlations implied by

estimated variances and covariances. With limited exceptions, regimes 1 and 2 are characterized by moderate

15Guidolin and Ono (2006) also show that the fit provided by a VAR(1) model is rather poor. For instance, the VAR does

match the data sample means, i.e. over the long run the VAR(1) forecasts values for the variables that are often radically different

from those observed on average in-sample.16We also obtain evidence of predictability of macro variables, especially IP growth which is not only persistent, but also

predicted by past T-bill yields and default spreads with significant coefficients, similarly to Stock and Watson (1989). Interestingly,

the ability of asset returns to predict macroeconomic conditions − principally future real growth − seems confined to single-regimemodels, see e.g. James, Koreisha, and Partch (1985). When k = 4 most coefficients lose significance. This means that the bulk of

predictability for inflation and real growth comes from the regime switching structure (and past macroeconomic conditions), and

not from financial markets.

11

Page 13: The Economic and Statistical Value of Forecast ... · 2For a special case, Aiolfi and Timmermann (2005) provide conditions under which the population MSFE of an equally-weighted

volatilities of the shocks and by correlations which tend to be smaller (in absolute value) than in the single-

state VAR(1) model of Panel A. On the opposite, regimes 3 and 4 imply higher volatilities and (at least for

a majority of pairs) larger correlations in absolute value. In fact, Table 2 and Figure 3 help us giving some

economic interpretation to the four regimes. Regime 1 is a bull/rebound state characterized by high equity

risk premia (14.5% on annualized basis), low or negative real short term interest rates, relatively high inflation

(4.6% on annual basis), and high dividend yields. In this regime, all variables display moderate volatility,

e.g. 13% for excess equity returns, 2.4% for excess bond returns, and 2.5% for inflation. This is a rebound

state because its persistence is moderate (approximately 10 months) and it tends to follow bear regimes: the

estimated transition matrix in Table 2 shows that starting from a bear/recession regime, 17% of the time the

system accesses a rebound (76% it stays in a bear regime). As a result, the mean dividend yield tends to be

exceptionally high (5.2% vs. a historical mean of 3.8%), an indication of the existence of good bargains in the

stock market. The exceptional stock market performance tends to be disjoint from real growth, which has

actually an unconditional mean of only 0.84% per annum. Consistently, the yield curve is flat (the annualized

term premium is 1.9%). Historically this regime coincides with the stock market bubble of 1927-1929, the

Great-Depression rebound of 1934-1937, and most of the WWII and immediate post-war years. After one

spike in the mid-1950s, the occurrences of this regime have been rather episodic, although some late periods

in the tech bubble of 1999-2000 are captured by this state.

Figure 3 approximately here

Regime 2 is a stable (low volatility) regime characterized by good real growth (2.9% per year) and moderate

inflation (3.2%). This a persistent regime (15 months on average) in which also equity risk premia are fairly

high (5.4%), although equity prices correspond to much higher multiples than in regime one (the dividend

yield has unconditional mean of 2.8%, below the historical sample mean). As experienced in the 1990s, real

short term rates are low and default-free credit cheap, just in excess of 1% per year. In fact, regime 2 captures

most of the booming years between the mid 1950s and 1974 (the interruptions correspond to officially dated

NBER recessions, picked up by state 4). The 1990s as well as the more recent, 2002-2004 period are entirely

captured by regime 2.

Regime 3 describes periods of intense real growth, the initial stages of the business cycle when the econ-

omy emerges from a trough. In fact, regime 3 is scarcely persistent (5 months on average). This state is

characterized by high real growth and equity premia (5.9%), and an upward sloping yield curve. Figure 3

shows that − among other periods − the early 1980s and 1990s are captured by regime 3. Finally, regime4 represents a bear/recession state, in which risk premia are small, dividend yields relatively high (as stock

prices decline), and inflation and real growth are both negative. Consistently with this interpretation, the

default spread is high in this state (25 basis points vs. a historical mean of 9 points only), while its duration is

moderate (4 months), which reflects the fact that recession and bear markets are generally short-lived. This

state also implies high volatility. Figure 3 shows that regime 4 picks up all major US recessions after WWII,

in addition to a long period that matches the so-called Great Depression.

Figure 4 reinforces the impression that the four-state model provides an accurate fit to the data by plotting

the in-sample fitted values for the eight series under examination. On the top of each panel, correlations

between fitted and observed values are reported. First, we notice that in many instances, the MSIH(4,0)-

VAR(1) values are simply much more volatile (hence able to track the underlying series) than VAR(1) fit,

which is to be expected. A careful eye may even detect the existence of periods (e.g., the 1990s) over which the

12

Page 14: The Economic and Statistical Value of Forecast ... · 2For a special case, Aiolfi and Timmermann (2005) provide conditions under which the population MSFE of an equally-weighted

behavior of the fit values across the two models is clearly heterogeneous. Second, the four-state model does

a superior job at matching the dynamics of most of the series: for six out of eight, the correlation between

actual values and fitted ones is higher under a MSIH(4,0)-VAR(1) than under a VAR(1).

Figure 4 approximately here

5. Forecasting Performance

As a first step towards investigating the value of forecast combinations under regime switching, in this Section

we proceed to establish that a four-state MSIH(4,0)-VAR(1) also precisely predicts most of the variables of

interest, especially excess asset returns. In this Section, a Reader will also notice that our focus gradually

shifts from the overall vector yt to xt, the set of excess stock and bond returns. The role of other variables —

especially macroeconomic predictors — never completely fades as they remain important forecasting variables

in the experiments that follow, but from this point onwards our attention is mostly devoted to asset returns

because they are the variables for which an assessment of the economic value of using forecast combinations

is most easily obtained.

5.1. Mean Square Forecast Error Results

When the models are flexible because of their rich parametrization, accuracy of fit is relatively unsurprising.

However, rich parametrizations are also well known to introduce large amount of estimation uncertainty

which may end up deteriorating their out-of-sample performance, see e.g., Rapach and Wohar (2006). In the

predictability literature this has been stressed among the others by Chan, Karceski, and Lakonishok (1998)

who − studying the out-of-sample predictability of stock returns − found that except for the term and defaultspreads, macroeconomic variables tend to perform poorly. Goyal and Welch (2003) report that the dividend

yield is a good predictor of stock returns only when the forecasting horizon is longer than 5 to 10 years.

Otherwise the fit is good only in-sample. Neely and Weller (2000) re-examine the findings of Bekaert and

Hodrick (1992) using a predictive metrics. They show that VAR models are outperformed by much simpler

benchmarks, like random walks. They suggest that the poor forecasting performance may be due to underlying

structural changes. Similarly, Bossaerts and Hillion (1999) argue that even the best linear models contain no

out-of-sample forecasting power even when the specification of the models is based on statistical criteria that

should penalize over-fitting. They speculate that the parameters of the selected models may be changing over

time so that the correct model ought to be nonlinear, possibly of a regime switching type.

To assess whether taking regimes into account leads to any improvement in prediction performance, we

implement the following ‘pseudo out-of-sample’ recursive strategy. For a given model, we obtain recursive

estimates over expanding data windows 1926:12 - 1985:01, 1926:12 - 1985:02, etc. up to 1926:12 - 2004:12.

This gives a sequence of 240 sets of parameter estimates specific to each of the models. For instance, the

regime switching model (3) generates 240 sets of regime-specific intercepts, covariance matrices, and transition

matrices, as well as regime-independent VAR(1) coefficients. At each final date in the expanding sample − i.e.on 1985:01, 1985:02, etc. up to 2004:12-h, where h is the forecast horizon (in months) — we calculate h-month

ahead forecasts for each of the 8 variables under study, i.e. including both financial and macroeconomic

variables. For concreteness, in what follows we focus on h = 1, 12 months. Calling y(M,h)t the forecast

generated at time t, by model M , at horizon h (in the following we drop the variable index, n = 1, ..., 8).

13

Page 15: The Economic and Statistical Value of Forecast ... · 2For a special case, Aiolfi and Timmermann (2005) provide conditions under which the population MSFE of an equally-weighted

Finally, we evaluate the accuracy of the resulting forecasts, by calculating the resulting forecast errors, e(M,h)t ≡

yt+h − y(M,h)t .

Table 3 reports summary statistics concerning the quality of relative forecasting performance. In particular,

we report three statistics illustrating predictive accuracy: the root mean-squared forecast error (RMSFE),

RMSFE(M,h) ≡

vuut 1

240− h

2004:12−hXt=1985:01

³yt+h − y

(M,h)t

´2,

the prediction bias

Bias(M,h) ≡ 1

240− h

2004:12−hXt=1985:01

³yt+h − y

(M,h)t

´,

and the standard deviation of forecast errors,

SD(M,h) ≡

vuut 1

240− h

2004:12−hXt=1985:01

"(yt+h − y

(M,h)t )− 1

240− h

2004:12−hXt=1985:01

³yt+h − y

(M,h)t

´#2.

Notice that the three statistics are not independent as it is well known that MSFE(M,h) ≡£Bias(M,h)

¤2+£

SD(M,h)¤2, i.e. the MSFE can be decomposed in the contribution of bias and forecast error variance. At a

one-month horizon, the four-state model seems to be rather useful for forecasting purposes, as it displays the

lowest MSFE for five out of eight variables. When compared to the other six benchmarks, the reduction in

MSFE is particularly important for excess stock returns and real money growth (the ratio of the four-state

MSFE to the next best MSFE are 0.88 and 0.82, respectively). However for two variables (excess bond returns

and the default spread) simple random walk benchmarks outperform all other models, although the difference

relative to the four-state model is small. Consistently with the results in Section 4.2, when h = 1 a richer

three-state model with regime-dependent VAR coefficients is systematically dominated by the four-state model

in terms of MSFE minimization; also, a VAR(1) seems too simple to be able to produce useful forecasts.17

Moreover, also Pesaran-Timmerman (PT) recursively updated regressions and the univariate AR models

provide disappointing performances, especially for what concerns excess asset returns and real money growth.

Finally, the random walk benchmarks produce biases that are close (or even lower) to those characterizing the

four-state model, the difference being in the variances of the forecast errors: regime switching models produce

higher biases (i.e. they often miss the actual point value of the forecast variable), but they seem to be able to

systematically move in a way that reduces the error variance. This is what one expects if the model’s regimes

well identify in real time the economy’s turning points.

Table 3 approximately here

At longer horizons, the superior performance of the four-state model tends to deteriorate slightly, while

the three-state model acquires merit. For instance, at h = 12 the MSIAH(3,1) presents the lowest MSFE for

half of the variables, although the four-state model still outperforms all benchmarks in predicting excess stock

returns, the dividend yield, and real money growth. The performance of PT-style recursive regressions remains

slightly inferior to either regime switching (e.g., excess returns and inflation), or falls very close to the accuracy

17The exceptions are the default spread, which is best modeled as a three-state process, and the inflation rate which is best

predicted by a simple VAR(1).

14

Page 16: The Economic and Statistical Value of Forecast ... · 2For a special case, Aiolfi and Timmermann (2005) provide conditions under which the population MSFE of an equally-weighted

of simpler models (VAR and random walks). Interestingly, the family of regime switching models comes to

consistently outperform single-state (as well as naive) benchmarks at longer horizon: this is the case for five

out of eight variables at h = 12. This is relatively unsurprising in the light of the evidence we have offered in

Section 4 of the fact that regime switching models generally perform well at fitting the multivariate density

of the data: as h grows, forecasts from models with regimes rely less and less on the accurate identification of

the most recent turning points of the economy, and increasingly on the overall properties of the unconditional

density of the data.

5.2. Testing Differential Predictive Accuracy

Further examination of Table 3 reveals that — although rankings match our general comments — differences

between models are often modest, which casts some doubts on how useful the four-state model may be in

practice. We therefore proceed to assess whether the out-of-sample performances are different enough to

allow us to draw any conclusions on the relative precisions of alternative models by implementing the forecast

accuracy comparison tests proposed by Diebold and Mariano (1995). For concreteness, we focus on the

forecasts of excess stock and bond returns only. Let

dif(m,n,h)t ≡

³e(m,h)t

´2−³e(n,h)t

´2be the differential loss (in our case the standard square loss) of model m relative to the loss from model n.

We can test the significance of the differences between two sets of forecast errors based on the statistic

DM(m,n)h =

1240−h

P2004:12−ht=1985:01 dif

(m,n,h)tbσ(dif (m,n,h)

t )

a∼ N(0, 1), (9)

where bσ(dif (m,n)t,h ) is an estimate of the standard error of the loss function differential, e.g. the HAC estimator

bσ(dif (m,n)t,h ) =

hXj=−h

Covhdif

(m,n)t,h , dif

(m,n)t+j,h

i.

Tables 4 and 5 illustrate the results of such tests. Table 4 concerns h = 1 and Table 5 h = 12. Moreover,

values of DM(m,n)h appearing below the main diagonal concern forecasts of excess stock returns, while those

above the main diagonal refer to forecasts of excess bond returns. A negative value of DM(m,n)h implies

that the model in row (m) outperforms the model in column (n) as it produces a lower sample MSFE loss.

Finally, in the tables we boldface values of DM(m,n)h which imply p-values of 0.05 or lower under Diebold and

Mariano’s (1995) standard asymptotic normal distribution. However — as discussed by West (1996) — caution

should be exercised when interpreting these results because the small sample distribution of the statistic in (9)

can show large departures from normality, thus complicating the task of conducting inferences on differential

predictive accuracy. We tackle the issues posed by possible departures of the small-sample distribution of

(9) from normality by block-bootstrapping the distribution of DM(m,n)h for all pairs of models and h = 1, 12

months. The simulation procedure consists of three steps:

1. Generate B bootstrap resamples of the sample indices 1, 2, ..., T where T is the length of the out-of-sample period. The block length L is set equal to the forecast horizon, h. This choice is justified by the

15

Page 17: The Economic and Statistical Value of Forecast ... · 2For a special case, Aiolfi and Timmermann (2005) provide conditions under which the population MSFE of an equally-weighted

overlapping nature of the realizations underlying the forecast error series.18 The bootstrap resampling

under a block length of L can be easily performed as follows:

(a) Determine the number of blocks v needed to span the entire sample size T, i.e. v ≡ int(T/L) + 1

where int(·) denotes the integer part of the rational number T/L.(b) For a = 1, ..., v draw τ ba as a discrete IID uniform over 1, 2, ..., T obtaining a collection of v time

indices τ b1, τ b2, ..., τ bv(c) The b-th resample indices are then given as τ b1, τ b1+1, ..., τ b1+L−1, τ b2, ..., τ b2+L−1, ..., τ bv, ..., τ bv+

L− 1. If it occurs that for some j ≥ 1 τ ba+ j > T, then set the time index to τ ba = τ ba+ j − T, i.e.

the resampling is restarted from the beginning of the sample in case the scheme attempts to draw

indices that exceed the overall sample size T.

(d) repeat a.-c. for b = 1, 2, ..., B.

2. For each of the B resamples of length T indexed by b we calculate (9), dDM(m,n)

h .

3. At this point a small-sample simulated distribution of DM(m,n)h is obtained , dDM

(m,n)

h,b Bb=1. We reportp-values for two-sided tests of the null of zero differential predictive accuracy, i.e. the percentage of the

bootstrapped simulations such that

|dDM(m,n)

h,b | > |DM(m,n)h |,

where DM(m,n)h corresponds to the sample value of the statistic.

Table 4 approximately here

In the bootstrap implementation in the paper we set B = 50, 000 while the overall sample size T clearly

depends on the forecast horizon h. Tables 4 and 5 report the bootstrapped p-values in parenthesis, below the

values for (9). Table 4 shows that over short horizons the MSIH(4,0)-VAR(1) provides exceptional performance

for excess stock returns, since the model outperforms all other forecasts in a statistical significant fashion. The

associated p-values are in fact below 0.01 under the asymptotic distribution for (9), while they generally fall

around 0.05 when the small sample distribution of the statistic is bootstrapped. The only exception concerns

the comparison to the random walk, where the asymptotic p-value exceeds 0.05 and the bootstrapped one

is relatively large (0.16). Interestingly, we uncover evidence that the richer three-state MSIAH(3,1) model

outperforms a few of the other forecasts, although in this case the bootstrapped p-values tend to fall around

0.1. When it comes to excess bond returns, Table 4 shows that only two conclusions can be reached. First, the

four-state model outperforms in a significant way PT’s predictive regressions and the AR models. Second, in

this case the random walk guarantees an interesting precision, for example significantly outperforming PT’s

methods and the three-state model.

Table 5 approximately here

18For instance, a negative return for some of the assets at time t that falls in the extreme left tail of the predictive density under

some given statistical model is likely to generate large and negative forecast errors from all 12-month ahead forecasts generated

between time t − 12 and t − 1. To avoid considering these sub-sequences of errors as an indication of structure in the forecasterror series, we set the block length to h.

16

Page 18: The Economic and Statistical Value of Forecast ... · 2For a special case, Aiolfi and Timmermann (2005) provide conditions under which the population MSFE of an equally-weighted

Table 5 presents results concerning 12-month forecasts. Importantly, results are qualitatively similar —

only marginally weaker in terms of bootstrapped p-values — than in Table 4. The four-state model keeps

outperforming most other competitors with some statistical strength when it comes to predict excess stock

returns. However the MSIAH(3,1) framework is now inferior to all other forecasts, which is consistent with

the idea that a three-state model might be not correctly capturing the shape of the ergodic density of returns

data. Interestingly, the performance of the four-state model improves (vs. h = 1) in terms of performance on

excess bond returns, since now the implied MSFE is significantly lower to the one produced by the three-state,

the VAR(1) and the random walk. The performance of the single-state VAR(1) model is particularly poor.

In summary, predictive accuracy tests show that differences in Table 3 could not entirely attributed to

chance. Consistently with much forecasting literature, at short horizons it remains true that outperforming

the random walk or simple constant expected returns benchmarks is difficult. At longer horizon, it is clear

that the flexibility of mixture (regime switching) models may be required to capture the salient features of

the multivariate distribution of financial and macroeconomic variables. However, careful scrutiny of Tables 4

and 5 also highlights that predictive accuracy comparisons are hardly unidirectional. We find a few instances

in which while model A fails to significantly outperform model B, and B fails to outperform C, the same

cannot be said about A outperforming C.19We take this as an indication that relative performances are rather

heterogeneous on different parts of the (pseudo) out-of-sample period. This opens the door to the possibility

that carefully combining forecasts produced by different frameworks may deliver important payoffs. This is

the issue that the next two Sections explore.

6. Are Forecast Combination Useful? Statistical Evidence

The simultaneous availability of forecasts from linear and non-linear, from univariate and multivariate, and

from models with constant vs. time-varying parameters represent one of the classical cases in which forecast

combinations are generally thought of as valuable: unless a particular forecasting model that (uniformly, i.e.

for each out-of-sample period) generates smaller forecast errors than all of its competitors can be identified,

forecasts combinations will always ex-ante reduce the overall MSFE because they offer “diversification gains”.

The practical value of this insight for our case study is analyzed in this Section.

6.1. The Forecast Combination Problem

We are interested in forecasting at time t the future values after h ≥ 1 months of the first two elements ofyt, i.e. excess stock and bond returns, xt+h ≡ [xst+h xbt+h]

0. Call xst,t+h and xbt,t+h two F × 1 vectors — one for

excess stock returns, the other for excess bond returns — that collect F distinct forecasts produced at time t.

Finally, call xst,t+h and xbt,t+h the forecast combinations obtained as

xst,t+h = gs(xst,t+h;ωst,t+h)

xst,t+h = gb(xbt,t+h;ωbt,t+h),

19For instance, for h = 12 and with reference to excess bond returns (panel above the main diagonal), both MSIH(4,0)-VAR(1)

(p-value is 0.15) and the random walk fail to outperform the AR model (p-value 0.10), suggesting that MSIH(4,0)-VAR(1) and

the random walk must be providing “similar” performances. However MSIH(4,0)-VAR(1) significantly reduces the MSFE vs. the

random walk (p-value is 0.01).

17

Page 19: The Economic and Statistical Value of Forecast ... · 2For a special case, Aiolfi and Timmermann (2005) provide conditions under which the population MSFE of an equally-weighted

where gs(·;ωst,t+h) and gb(·;ωb

t,t+h) are functions (RF → R) — parameterized by two (F + 1) × 1 vector ofweights ωs

t,t+h and ωbt,t+h — that combine (pool) the F forecasts into one aggregate forecast. For instance,

gs(·;ωst,t+h) and gb(·;ωb

t,t+h) might be restricted to be linear affine functions, in which case

g[·](x[·]t,t+h;ω[·]t,t+h) = ω0,t,t+h +

FXf=1

ωf,t,t+hx[·]t,t+h, (10)

where the subscript [·] refers to either stocks or bonds.20 For instance, when an equal weighting scheme isimposed, then ωf,t,t+h = 1/F, f = 1, ..., F and ω0,t,t+h = 0.

In general, while the structure of the functions g[·](·;ωst,t+h) tends to imposed by the researcher, the

combination weights are not exogenously fixed, but they are instead computed using the principle that decision

makers do not attach any intrinsic value on forecasts: it is instead the forecasts (or combination thereof) that

minimize their overall loss function to matter. In particular, we follow standard practice and assume that the

loss function only depends on the forecast error, est,t+h ≡ xst,t+h − xst,t+h, ebt,t+h ≡ xbt,t+h − xbt,t+h, and that the

loss function is of a standard, symmetric quadratic type:

L³e[·]t,t+h

´≡³e[·]t,t+h

´2.

Therefore the forecast combination problem becomes

minω[·]t,t+h

Ehx[·]t,t+h − g[·](x[·]t,t+h;ω

[·]t,t+h)

2¯x[·]t,t+h

i, (11)

i.e. the forecast combination weights are chosen to minimize the (conditional) MSFE. In this case, it is

straightforward to show that

ωst,t+h Ä g[·](x

[·]t,t+h; ω

st,t+h) = E

hx[·]t,t+h

¯x[·]t,t+h

i, (12)

i.e. the weights will be chosen to allow g[·](x[·]t,t+h; ωst,t+h) to best approximate the conditional expectation of

the excess asset returns given the information contained in the F × 1 vector of forecasts. For concreteness,and following in the steps of Bates and Granger (1969), we will restrict our attention to the case in which

g[·](x[·]t,t+h;ω[·]t,t+h) is a linear function, as in (10). This choice together with (12) implies that ω

st,t+h can be

simply estimated as a vector of least squares estimates in the regression:

x[·]t,t+h = ω0,t,t+h +

FXf=1

ωf,t,t+hx[·]t,t+h + η

[·]t,t+h. (13)

As shown by Chong and Hendry (1986), under squared loss function (13) also provides a test of pair-wise

encompassing. For instance, when F = 2 ω0,t+h = ω2,t+h = 0, ω1,t+h = 1 implies that forecast 1 encompasses

forecast 2, while ω0,t+h = ω1,t+h = 0, ω2,t+h = 1 brings to the conclusion that forecast 2 encompasses 1. When

one forecast function encompasses another, this means that combining the two forecasts will not reduce the

overall MSFE. However, an indication that for some of the forecasts in (13) the corresponding weight is nil,

will automatically lead to avoid using the forecast combination, in the sense that (trivially) g[·](x[·]t,t+h; [0 1

0]0]) = x1t,t+h and g[·](x[·]t,t+h; [0 0 1]0]) = x2t,t+h.

20Granger and Ramanathan (1984) recommend adding a constant intercept under squared loss; Elliott and Timmermann (2004)

reach the same conclusion for a variety of loss functions.

18

Page 20: The Economic and Statistical Value of Forecast ... · 2For a special case, Aiolfi and Timmermann (2005) provide conditions under which the population MSFE of an equally-weighted

Because

E

⎡⎣⎧⎨⎩x[·]t,t+h − ω0,t,t+h −

FXf=1

ωf,t,t+hx[·]t,t+h

⎫⎬⎭2 ¯¯ x[·]t,t+h

⎤⎦ = V arhe[·]t,t+h

¯x[·]t,t+h

i+nEhe[·]t,t+h

¯x[·]t,t+h

io2,

(e[·]t,t+h is defined as the forecast combination error) problem (11) implies that each of forecast functions entering

x[·]t,t+h is useful in the measure in which they help predicting either the (conditional) bias of the pooled forecast,

or its conditional variance. As stressed by Timmermann (2005), it is common to find that E[e[·]t,t+h|x

[·]t,t+h]

exceeds the bias produced by the best forecasting model in the vector x[·]t,t+h; however, in practice forecast

combinations often reduce the MSFE by reducing V ar[e[·]t,t+h|x

[·]t,t+h] by substantial amounts.

Besides equal-weighting schemes — i.e. ωf,t,t+h = 1/F, f = 1, ..., F and ω0,t,t+h = 0 — another popular

combination benchmark suggested by Winkler and Makridakis (1983) and Stock and Watson (2001) consists of

computing weights that simply reflect the performance of each individual model relative to the performance of

the average model but ignore correlations across forecasts. As first remarked by Newbold and Granger (1974),

in small samples and with large numbers of forecast functions F , correlation may prove hard to estimate with

any precision. Letting MSFEf,t,t+h ≡Pt

τ=1(xτ − xf,τ ,τ+h)2 the mean-squared error from model f = 1, ..., F,

Stock and Watson suggest to set

ωf,t,t+h =[1/(MSFEf,t,t+h)

α]PFf=1[1/(MSFEf,t,t+h)α]

(14)

and find that the scheme works well in practice. In the following we use equation (14) setting α = 1, i.e. we

experiment with a scheme that assign the higher weight the lower is the MSFE.

Finally, an important issue often debated in the forecast combination literature is whether any constraints

ought to be imposed on the weights ωt,t+h, chiefly whether it makes sense to either impose that ω0t,t+hιF = 1

(i.e. the combination weights must sum to one) or that 0 ≤ ωf,t,t+h ≤ 1 f = 1, ..., F (i.e. to rule out that the

combined forecasts may lie outside the range of individual forecasts). On one hand, rigorous arguments have

been made in favor of both restrictions: when the F forecasts are all unbiased, ω0t,t+hιF = 1 guarantees that the

forecast combination will also be unbiased; under the assumption of unbiasedness, Diebold (1988) argues that

ω0t,t+hιF = 1 minimizes the chances that the forecast errors from the combination may be serially correlated

and hence predictable (viz. inefficient). The convexity restriction 0 ≤ ωf,t,t+h ≤ 1 avoids the possibility that— given a range of possible h−step ahead forecasts at time t — the combined forecast may lie outside the

scope of the originally predicted values. On the other hand, the empirical results to follow will offer plenty of

evidence of the presence of biases in the forecasts. This is rather intuitive when facing structural shifts in the

intercept of the process followed by the variables of interested, as documented in Section 4.2. Additionally,

and similarly to Bunn (1985) and Guidolin and Timmermann (2005b), it is important to recognize that the

fact that a combination weight is negative does not imply it has no value and negative combination weights

can and do help providing more precise forecasts than the original input forecast functions.21

21Tests of both restrictions led to strong rejections for all the least-squares combination regressions that follow.

19

Page 21: The Economic and Statistical Value of Forecast ... · 2For a special case, Aiolfi and Timmermann (2005) provide conditions under which the population MSFE of an equally-weighted

6.2. Recursive Estimates of Combination Weights

We start by computing recursive combination weights over our (pseudo) out-of-sample period, i.e. over

sequences of expanding windows of data 1985:01 - 1985:12, 1985:01 - 1986:01, etc. up to 1985:01 - 2004:12−h.22

For a variety of model combinations, we do that in two ways:

• By estimating the weights from regressions of type (13) and in which F is set to either two (i.e. combi-

nations for pairs of models) or to seven, which is the total number of forecasting models introduced in

Section 2. In the following we call these weights “OLS weights”.

• By setting the weights according to equation (14), when α = 1 and F = 7. In the following we call theseweights “inverse-MSFE weights”.

As anticipated, we also compute equally weighted forecast combinations with F = 7. It must be noticed

that pairwise OLS weights are also computed for the case in which two different regime switching forecasts

— the ones from the MSIH(4,0)-VAR(1) and those from the MSIAH(3,1) — are combined. This is a very

interesting experiment that might provide additional prediction accuracy in the case in which three- and

four-state models pick up structural shifts at different frequencies or reflecting heterogeneous economic forces.

Using the combination weights obtained in these three alternative ways, we proceed to quantify predictive

performance along lines similar to Section 5.

Figure 5 provides a few examples of the behavior of the recursively estimated weights. The first two plots

in panel A concern OLS weights for one specific pairwise combination, the one between the four-state model

and PT-type forecasts, when the predictors are selected by recursive maximization of the R2.23 Especially in

the case of excess stock returns, Figure 5 displays enormous variability in the estimated combination weights.

In particular, while the weight assigned to the regime switching forecasts hovers in the narrow range 0.55-0.58,

the weight received by the PT forecasts is highly variable, oscillating from very high values in excess of 14 to

the negative range. Even the intercept correction, although seemingly negligible in the scale of the plots, jiggles

a lot in the range -0.48 to 0.02. Clearly, such wild gyrations in combination weights produce large amounts

of parameter uncertainty that are likely to degrade the out-of sample performance of the combinations.

Figure 5 approximately here

Panel B of Figure 5 shows instead the inverse-MSFE weights for both excess stock and bond returns.

Consistently with our comments to Table 3, in the case of excess stock returns the four state model sys-

tematically receives a large weight (roughly 24%), followed by the three state model (around 15%), which

however gradually declines over time. The remaining 60% is allocated in approximately equal fashion across

the remaining five models, with the sigle-state VAR(1) receiving a slightly lower weight of 10-11%. In the case

of excess bond returns, weights are much more balanced and stable over time, with approximately 17% going

to the VAR(1) and the random walk, 15-16% each to the two regime switching forecasts, and the remaining

third to PT and AR univariate forecast functions. Clearly, inverse-MSFE weights imply substantially lower

parameter uncertainty.

22To avoid excess instability in the coefficients in the first part of the recursive experiment, we use the first 12 values of forecast

functions and prediction errors to provide the 12 initial values of the optimal weight series (not shown in the figures).23Additional plots concerning other pairwise combinations or 12-step ahead combination weights appear qualitatively similar

and are not reported to save space. They are available upon request.

20

Page 22: The Economic and Statistical Value of Forecast ... · 2For a special case, Aiolfi and Timmermann (2005) provide conditions under which the population MSFE of an equally-weighted

6.3. Statistical Tests

Tables 6 and 7 provide information on the statistical performance of forecast combinations for h = 1 and 12

months, respectively. In both tables, panels A1 show the results for pairwise forecast combinations “centered”

around the MSIH(4,0)-VAR(1) model, i.e. in which f = 1 is held fixed at the four-state forecast combinations.

Panels B1 report on similar experiments when the combinations are “centered” around the PT forecasts when

the predictors are recursively selected by maximization of the R2. We adopt this additional benchmark as

a way of a robustness check; we select this PT-style model because of its slightly better performance (for

h = 1) vs. similar univariate prediction models in Table 3. Finally, in both Tables 6 and 7, panels A2 and

B2 show results for a few benchmarks which do not imply selection of forecasts to be combined: when all

forecasts are combined by computing OLS weights (F = 7); equally-weighted forecast; inverse-MSFE weights.

Additionally, for reference, we copy the performance statistics for our favorite regime switching framework,

the four-state model.

For the 1-month horizon, Table 6 shows that forecast combinations ”centered” around the four-state model

fail to improve the MSFE. On the opposite, they increase the MSFE vs. just using the MSIH(4,0)-VAR(1)

model, from 3.8% to approximately 6.1% for stocks, and from 1.3% to 2.8% in the case of bonds. Interestingly,

most of the deterioration is attributable to an increase in the variance of the forecast errors. This is likely

due to the enormous variability in the estimated combination weights and in the resulting large parameter

uncertainty, see Figure 5. Panel B shows similar results for OLS weights centered on PT forecasts. In fact, in

this case there is even evidence that combining PT forecasts with regime-switching based predictions delivers

results which are sensibly inferior to all other sorts of combinations. Results hardly improve when F stops

being forced to be equal to two and all of the forecast functions are part of the combination.

Table 6 approximately here

The fact that forecast combinations may be susceptible to failure in the presence of large estimation errors

in the combination weights has been recently discussed by Yang (2004). Finite-sample errors in the estimates

of the combination weights can lead to poor performance of combination schemes that can be proven to

dominate in large samples, in part because of the prevalence of strong multicollinearity among the forecasts

that are being combined. In the same way in which non-stationarities and structural change may provide a

justification for the use of forecast combinations, they can also lead to instabilities in the combination weights

and deteriorate the out-of-sample performance of the combination, see e.g., Palm and Zellner (1992).

However panels A2 and B2 of Table 6 show that when parameter uncertainty related to estimation of the

combination weights is wiped out — by using an equal weighting scheme — pooling forecasts delivers interesting

improvements in forecast accuracy. The RMSFE declines to 2.2% for excess stock returns and to 1.1% in

the case of excess bond returns; both values are lower than the best among pairwise combinations and the

four-state model. The improvement is almost entirely explained by a reduction in the forecast error variance,

which is again consistent with the idea that in panels A1 and B1 the high variability of the OLS weights

is what causes the failure of pairwise pooling. This result echoes previous studies which have found that

the simple weighting schemes are generally hard to beat. The reasons of this fact are discussed in Dunis,

Laws, and Chauvin (2001) and Hendry and Clements (2002). The most convincing explanation is that the

estimation uncertainty of the optimal weights offsets the advantage of using the forecast combination.

When estimation uncertainty is ”intelligently” dealt with — by setting the combination weights to be

21

Page 23: The Economic and Statistical Value of Forecast ... · 2For a special case, Aiolfi and Timmermann (2005) provide conditions under which the population MSFE of an equally-weighted

relative performance indicators — the improvement is substantial and across the board, i.e. it involves both

excess stock and bond returns. Under inverse-MSFE weights, the RMSFE error declines to 1.6% for excess

stock returns and to 0.7% for bonds. These values are relatively small and certainly dwarfed by the magnitude

of the volatility of both series (5.5 and 1.9 percent, respectively).

Table 6 shows that in the presence of structural change/regimes, forecast combinations improve prediction

accuracy, although this result may be severely limited by the presence of pervasive estimation uncertainty

when the weights are estimated, for instance using regression schemes a’ la Granger and Bates (1969). In

fact, the MSFE improvement from using inverse-MSFE weights in forecast combinations turns out to be

statistically significant: the corresponding Diebold-Mariano (DM) test for differences in MSFEs between the

pooled prediction and the four-state model produces a p-value of essentially 0.00 (the DM-stat is 7.37) for

excess stock returns and again of 0.00 (the DM stat is 9.34) for excess bond returns.24 Similarly, when the

baseline model is the PT scheme, the implied p-values for a test of the null that equal-weights and inverse-

MSFE weights combinations predict as well as PT are around 0.01, with DM-stats of around 5 for excess

stock returns and of 10 for excess bond returns (in the latter case the p-value is essentially nil).

Finally, Table 7 extends these tests to a 12-month forecast horizon. Panels A1 and B1 show that the

results in Table 6 also apply to 1-year forecasts, although in this case the improvement brought about by the

combinations is more modest, e.g. from 4.6% for the four-state model to 4.1% in the case of stocks, and from

2.1% to 1.9% in the case of bonds. Once more, inverse-MSFE combinations turn out to be the best, while

combinations (even relatively tightly parameterized ones, with F = 2) based on OLS weights are not very

successful. However such improvements are no longer statistical significant. When the baseline model is the

four-state regime switching (panel A1), even the inverse-MSFE weights imply a p-value of 0.23 (the DM stat

is 1.81) for stocks and a p-value of 0.14 (the DM stat is 1.58) for bonds. In conclusion, forecast combinations

work well with short investment horizons and when estimation of the weights can be somehow constrained or

even avoided altogether.

Table 7 approximately here

7. The Economic Value of Forecast Combinations: Portfolio Implications

The finding that forecast combinations may lower the RMSFE for excess stock and bond returns fails to pin

down the more interesting issue of their actual economic value to decision makers. In fact, although many

decision makers might have an interest for the results so far and the general issue of whether the dynamic

linkages between financial markets and macroeconomic factors might have been subject to regime shifts, there

is a class of economic agents that has a straightforward use for the results in Tables 6 and 7: portfolio

managers. In other words, the squared loss function we have adopted so far has a lot of convenience to it, but

need not represent a realistic loss to decision makers. It seems that portfolio returns, their Sharpe ratio, or

some utility function of realized wealth may represent a much more informative criterion to decide whether

forecast combinations have any value.

In fact, money managers have a keen interest in exploiting statistical predictability to improve the return-

risk properties of their portfolios. In this sense, such decision makers would be mostly interested not in point

24The same applies to equally weighted forecasts: the p-value for a test comparing to the four-state model is 0.00 for both

excess stock and bond returns, with DM stats of 7.12 and 9.86. These p-values and the ones mentioned in the rest of the Section

were bootstrapped using the same algorithm discussed in Section 5.2.

22

Page 24: The Economic and Statistical Value of Forecast ... · 2For a special case, Aiolfi and Timmermann (2005) provide conditions under which the population MSFE of an equally-weighted

forecasts of future asset returns or in the ability of (3) to approximate their conditional joint density, but in

correctly forecasting (one-step ahead) Sharpe ratios,

SR(M,i)t,t+1 ≡

E(M)t [yit+1]q

V ar(M)t [yit+1]

,

where i indexes excess stock and bond returns, respectively.

Figure 6 shows 1-month ahead, predicted Sharpe ratios for both stocks and bonds. Such predictions

are calculated under each of the seven models introduced in Section 2. As in Sections 5 and 6, the ratios

are computed recursively, by updating parameter estimates over the expanding windows 1926:12 - 1985:01,

1926:12 - 1985:02, etc. up to 1926:12 - 2004:12.25 Although most models imply wild variation in the one-

month ahead prediction of the Sharpe ratios, a few qualitative facts can be noticed. First, as one would

expect, models with regimes imply wide swings in SRt,t+1, although a structural difference seems to exist

between MSIH(4,0)-VAR(1) and MSIAH(3,1), in the sense that the latter model implies sharper corrections

and a rather implausible tendency for SRt,t+1 to be predicted in the negative range. In fact, while the four-

state model implies a sensible sample mean of 0.10 for the equity Sharpe ratio (0.02 for bonds), the same

cannot be said for the three-state model, where the sample mean is negative (-0.02, the same number obtains

for bonds), which is inconsistent with equilibrium asset pricing principles.26 Second, as already noticed in

a related experiment by Guidolin and Ono (2006), the single-state model produces Sharpe ratios that are

systematically too low for both asset classes, and systematically lower for stocks than for bonds, which is also

counter-intuitive. Third, regression-based models predict values of SRt,t+1 for equities which are essentially

constant over time and that seem to fall at average levels which depart from the historical values commonly

used in finance, e.g. in excess of 0.25 per month vs. a historical level of approximately 0.11. Notice that

flat predictions for the Sharpe ratio are hardly compatible with active portfolio management. Fourth, the

four-state model gives reasonable risk-return insights. The equity Sharpe ratio fluctuates over time in a

counter-cyclical manner, i.e. SR(M,s)t,t+1 is high during recessions and declines during economic booms. Under

regime switching, the bond SR(M,b)t,t+1 is stable but still provides strong and possibly useful signals.

Figure 6 approximately here

7.1. Recursive Portfolio Weights

In this section we proceed to the systematic, recursive calculation of optimal mean-variance portfolio weights

and assess the comparative out-of-sample portfolio performance of (the best performing) forecast combinations

vs. the models introduced in Section 2. Assume an investor endowed with initial unit wealth has preferences

described by a simple mean-variance functional:

Vt = Et[Wt+1]−1

2λV art[Wt+1]

Wt+1 = wst (1 + xst+1 + rt) + wb

t (1 + xbt+1 + rt) + (1− wst − wb

t )(1 + rt), (15)

25Notice the existence of a structural difference between the regime switching and the remaining models, because while the

former explicitly imply time-variation in variances as well as risk premia, the regression-based (e.g. PT schemes, and univariate

AR) and random walk models allow variation in variances only as a result of the recursive updating of the estimates.26Interestingly, the four-state model produces sample means for the Sharpe ratios which are virtually identical to the random

walk model (no predictability in risk premia), despite the variability in the ratios themselves.

23

Page 25: The Economic and Statistical Value of Forecast ... · 2For a special case, Aiolfi and Timmermann (2005) provide conditions under which the population MSFE of an equally-weighted

where λ is interpreted as coefficient of risk aversion that trades-off (conditional) predicted mean and variance

of the one-step ahead wealth, and rt is a short-term (1-month) interest rate.27 At each time t in the sample, the

investor maximizes Vt by selecting weights wt ≡ [wst wb

t ]0 when the predicted moments are calculated using

some reference statistical model, e.g. our four-state model or some forecast combination scheme. Simple

algebra shows that:

wt =1

λdV art[xt+1]−1Et[xt+1],

Portfolio weights are calculated recursively using the recursive parameter estimates also employed in Sections

5 and 6.28 Problem (15) is solved when investors are prevented from selling any securities short, i.e. such that

w0ten ∈ [0, 1] ∀t and n = s, b.29 Of course, (15) corresponds to a very simplistic asset allocation strategy in

which preferences (absolute risk aversion) are assumed to be constant, and investors care only about mean and

variance. Additionally, such investors would have to be strictly myopic and ignore time-varying investment

opportunities. Notice that regime switching models imply rich (as well as dynamic, as the state probabilities

are recursively updated) departures from standard zero skewness and constant kurtosis levels, which may

be at odds with the assumption of mean-variance preferences, especially for horizons exceeding one month.

However, this seems to be a relatively straightforward way for us to collect some evidence bearing on the issue

of whether forecast combinations have any economic value for a portfolio manager.30

Table 8 shows a few statistics for recursive optimal portfolio weights obtained under four alternative

assumptions on the risk aversion coefficient, λ = 0.2, 0.5, 1, and 2.31 The table shows that different mod-

els/forecast combination schemes imply quite different average allocations to the different assets, as well as

different variability of the optimal portfolio weights. For instance, assuming λ = 1, the percentage invested

in stocks ranges from 100 under the random walk or PT-style predictions (both implying high predictions of

the equity Sharpe ratios) to 17 under a VAR(1) model; the weight assigned to bonds similarly ranges from

0 to 43 percent. Interestingly, regime switching models and forecast combination schemes end up implying

comparable portfolio weights, of approximately 60% in stocks, 20% in bonds, and 20% in cash. Notice that

this result is far from trivial because the combination schemes used in Table 8 involve with non-negligible

weights all the prediction models, as shown in Section 6.1. Finally, regime switching-based portfolio strategies

imply a structurally higher volatility of the optimal portfolio weights. At this point the key question becomes:

which portfolio allocation strategy pays off the most over a long sample such as ours?

Table 8 approximately here

27Notice that predictions of the interest rate fail to enter the portfolio problem because the (nominal) yield of 1-month T-bills

is known ex-ante. However, in our experiments we recursively update rt which therefore preserves a time index.28For instance, w1985:01 is based on estimated parameters obtained using data for the interval 1926:01 - 1985:01, w1985:02 on

estimates for 1926:02 - 1985:02, etc.29When short-sales are restricted, wt has no closed-form solution and is calculated numerically (by grid search).30Ang and Bekaert (2002a) and Guidolin and Timmermann (2005a, 2006a) explore the optimal asset allocation implications of

regimes and predictability in the joint (and time-varying) distribution of excess asset returns.31While λ = 0.2 is admittedly a rather low risk aversion coefficient, notice that introspection suggests that λ = 2 represents a

rather high level already. For instance, it is well known that in simple mean-variance asset allocation frameworks, the optimal

weight of a single risky asset is wt,t+1 = SRt,t+1/λ.At this point, the Sharpe ratio on the value-weighted CRSP portfolio over

the period 1926-2004 is 0.12; λ = 2 implies then a weight of 0.12/2 = 0.06, i.e. only 6% in US stocks. That represents a rather

conservative portfolio position. In fact the range for λ commonly spanned in the literature (see e.g. Jorion, 1986) is 0 - 2.

24

Page 26: The Economic and Statistical Value of Forecast ... · 2For a special case, Aiolfi and Timmermann (2005) provide conditions under which the population MSFE of an equally-weighted

7.2. Out-of-Sample Portfolio Performance

Table 9 completes our “economic” analysis by computing (pseudo) out-of-sample, one month portfolio per-

formance under different levels of λ and for the seven competing models/allocation strategies. Notice that

models 6 and 7 now identify forecast combination-driven strategies, based on either equal forecast weights

or on the inverse-MSFE weights introduced in Section 6. Additionally, we consider two benchmarks which

(albeit rather naive, as they ignore all sample information on the dynamics of asset returns) are often used

in practice: an equally weighted portfolio (“1/N”, see De Miguel et al., 2006) in which 1/3 is invested in

each of the available assets at each point in time; a similar “50-50” portfolio in which 50 percent shares are

allocated to stocks and bonds, respectively, at each point in time. The logic of introducing these two further

benchmarks is that they may end up yielding interesting performances for two reasons: First, the benchmarks

are completed unaffected by parameter uncertainty, since they fail to rely on any econometric framework

implying estimation issues. Second (see Section 7.4 for further details), when held over a one-month period

only, these portfolios imply no need for rebalancing and thefore completely escape the incidence of transaction

costs.32

Table 9 approximately here

In particular, we report means and medians of one-month (gross) portfolio returns, the lower and upper

values of a standard 90% confidence interval (that reflects the volatility of realized portfolio performance over

1985:01 - 2004:11), and the implied Sharpe ratio that adjusts mean returns to account for risk. The last

column of Table 9 also lists the realized, recursive mean-variance statistic for each model/strategy, i.e.

1

239

2004:11Xt=1985:01

ÃW(m)t+1

W(m)t

!− 12λ1

239

2004:11Xt=1985:01

("ÃW(m)t+1

W(m)t

!−

2004:11Xt=1985:01

ÃW(m)t+1

W(m)t

!#)2, (16)

where Wt+1/Wt is simply the realized gross portfolio return over a one-month holding period when model m

is employed. The table reports in bold the maximum values of mean and median portfolio returns, of the

Sharpe ratio, as well as of the mean-variance statistic across models.33

Table 9 reveals that forecast combinations outperform all other models in all cases and under all dimensions

— mean and median portfolio returns, Sharpe ratio, and a mean-variance objective. In particular, the perfor-

mance indicators are systematically best for the inverse-MSFE weighting scheme of Stock and Watson (2003),

although also equal weights tend to perform well. For instance, assuming λ = 1, inverse-MSFE combination

weights generate mean (median) monthly portfolio returns of 1.7% (1.6%), and a remarkable Sharpe ratio of

0.36. Even though the 90% (empirical) confidence band for realized portfolio returns remains rather wide —

spanning the interval -2.3% to 5.2% — this seems a quite remarkable performance. For comparison, a statistic

random-walk based strategy would yield a mean monthly return of 1.2% only, a Sharpe ratio of 0.20, and

imply an even wider 90% confidence band of [-4.5%, 5.3%]. De Miguel et al.’s (2006) “1/N” equally weighted

portfolio gives a mean portfolio return of 0.8% per month and considerably lower Sharpe ratio, 0.16. Among

32We thank an anonymous referee for pointing out the usefulness of these further benchmarks. Notice that over horizons longer

than one month, even simpler benchmarks such as “1/N” and “50-50” would actually imply the need to rebalance to take into

account that as assets give different returns, their portfolio weights change accordingly. Clearly, “1/N” and “50-50” differ because

the latter only invests in risky assets.33Clearly, the Sharpe ratio and the mean-variance statistic are strictly related. The difference is the latter is more tightly

connected with the objective function that the vector of optimal weights is maximizing.

25

Page 27: The Economic and Statistical Value of Forecast ... · 2For a special case, Aiolfi and Timmermann (2005) provide conditions under which the population MSFE of an equally-weighted

the strategies that do not exploit combinations, the four-state model comes up on top, with a mean portfolio

return of 1.1% that however represents satisfactory compensation for risk (the Sharpe ratio is 0.21) since

realized returns are much less volatile when regimes are taken into account (e.g., the 90% confidence band is

[-3.9%, 4.6%]). Forecast combinations also systematically maximize the sample analog of the mean-variance

objective, equation (16).34

Although a difference of 0.5% in realized portfolio returns between the random walk benchmark and

forecast combination-based strategies may appear small (but the difference in realized Sharpe ratio is rather

large), there is in fact strong statistical evidence that a forecast combination scheme may produce superior

portfolio outcomes than any other models, including those that take into account the presence of regimes.

For instance, assuming λ = 1 and treating portfolio returns as the relevant loss function, a DM-type test

of the null of equal portfolio performance between the inverse-MSFE scheme and the MSIH(4,0)-VAR(1)

produces a p-value of 0.03.35 When the relevant loss function is identified — probably more correctly — with

the mean-variance statistic, the p-value is instead 0.001.36 In short, with a small margin of type I error (at

worst 6.7%), both portfolio returns and returns adjusted for variability are significantly higher under the best

forecast combination schemes than under the best fitting and predicting models that capture regimes.37

What is then the price tag we may attach to the possibility of supplementing careful modeling of regime

switches in predictability relationships involving US excess asset returns with forecast combination schemes?

Our results show that — depending on the risk aversion coefficient — such a value ranges from a 8.4% increase

in mean annualized portfolio returns to an increase in the realized Sharpe ratio from 0.20 to 0.36. Both

measures seems hardly negligible and may represent a powerful incentive for money managers to seriously

entertain pooling the predictions yielded by a number of competing models as a serious alternative to simply

selecting whichever of the models may have guaranteed a superior performance in the past.

7.3. Sub-Sample Performance

On an intuitive dimension, regime switching models may improve out-of-sample performance — both at fore-

casting asset returns and in realized portfolio terms — because they allow for different estimated models of the

relationship between asset returns and predictors to characterize periods of bull and bear markets, see e.g.

the discussion in Guidolin and Ono (2006). If this is the case, we should find that the performance of either

regime switching models or models that pool forecasts produced by models that account for regime shifts

should not be too sensitive to specific sub-periods of the overall (pseudo) out-of-sample interval, 1985-2004,

provided that the periods are long enough to incorporate both bull and bear markets. Therefore we have

divided the 20 years used in our experiment in four sub-samples, each including 5 years and consequently an

adequate number of regime shifts (see Figure 3), and examined the performance of each of our nine models

34For comparison we have also computed out-of-sample performance statistics for the 100% stocks and bonds portfolios, re-

spectively. The 100% stock portfolio gives a mean monthly return of 1.2% (i.e. the implied risk premium is 5.6% per year) and

a Sharpe ratio of 0.20. The corresponding statistics for the 100% bond portfolio are 0.8% (the implied risk premium is 0.9% per

year) and 0.11. In our sample, diversification obviously pays off on the basis of most econometric frameworks.35The p-values are boostrapped using the scheme introduced in Section 5.2. The p-values for this null hypothesis obtained in

the cases of λ = 0.2 and 2 are 0.02 and 0.07, respectively.36The p-values for this null hypothesis obtained in the cases of λ = 0.2 and 2 are 0.00 and 0.003, respectively.37It is instead more difficult to distinguish the two combination schemes — equally weighted and inverse-MSFE relative weights —

in the sense that portfolio returns do not statistically exceed the ones from equal weights. However the null of equal mean-variance

objectives may again be rejected for all levels of risk aversion.

26

Page 28: The Economic and Statistical Value of Forecast ... · 2For a special case, Aiolfi and Timmermann (2005) provide conditions under which the population MSFE of an equally-weighted

along the usual dimensions. To save space, we report only results for the two most plausible risk aversion

coefficients, λ = 0.5 and 1.

Table 10 reports the relevant results. Importantly, even though the specific performance measures prove

to be highly time-varying (in particular, as one should expect, performances are excellent over 1995-1999

and rather disappointing for the periods 2000-2004), the relative ranking across models remains essentially

identical to Table 9 for each of the four 5-year periods. The forecast combination-based portfolio strategies

turn out to be rather special because they are the only ones to always generate positive Sharpe ratios, although

the ratios vary from the 2000-2004 minimum of 0.07-0.09 (for inverse-MSFE portfolios) to 0.43-0.62 in the

booming 1995-1999 years.38

Table 10 approximately here

7.4. Transaction Costs

One further dimension in which alternative investment strategies may differ is for their implication for the

portfolio turnover, i.e. the relative frequency (and size) at which portfolio weights are changed over time,

as a result of model-specific revisions of expected returns, variances, and covariances. Clearly, turnover may

actually affect out-of-sample portfolio performance when there are non-negligible transaction costs which

penalize the decision to change portfolio structures. In this respect, while a few of our benchmarks (such

as the “1/N” and “50-50” portfolios) imply zero turnover and therefore no transaction costs, a few of the

other models (especially, regime switching models and forecast combination-based portfolios which also involve

regime switching forecasts) may in principle be quite expensive, as they require active portfolio rebalancing.

We therefore proceed to recompute monthly means and medians of portfolio returns, Sharpe ratios, and

realized mean-variance utility in the presence of transaction costs, tc.We specify the transaction cost function

after Lynch and Balduzzi (2000):

tct = φp

3Xj=1

|wjt − wj

t−1|+ φfImaxj |wjt−wjt−1|6=0

,

where wjt is the optimal portfolio weight of asset j at time t, Imaxj |wjt−w

jt−1|6=0

is an indicator function that

takes unit value when any of the differences wjt − wj

t−1 is non-zero (i.e., the weight of asset j has changed

between month t− 1 and t). φp and φf are two parameters which measure the unit incidence of transaction

costs. In what follows we specify φp = 0.25% and φf = 0.1%.39 In essence, instead of computing and reporting

Wt+1, we focus on Wt+1 − tct+1.

Table 11 shows the results for the two most realistic values of λ, 0.5 and 1. The realized out-of-sample

performance measures ought to be compared to the ones in Table 9. As one should expect, the statistics for

38The only minor difference (of no relevance to risk-averse investors) is that periods can be found for which either Neeley-Weller’s

or Pesaran-Timmermann’s models yield the highest mean portfolio returns.39These are the baseline values used in Lynch and Balduzzi (2000, pp. 2296-2297). The fixed cost parameter translates into

paying a fee of $10 whenever a $100,000 portfolio is reshuffled. The proportional cost parameter implies a roundtrip cost of 0.5

percent. For instance, the Fidelity’s Spartan Total Market Index Fund is an index fund that attempts to track the value of a U.S.

value-weighted equity portfolio and it has been charging a redemption fee of 0.5 percent on fund shares sold within three and six

months of purchase, respectively. Although it is possible to think that equity portfolios may implier higher values for φp and φf ,

the opposite applies to bond and especially cash portfolios (which are almost free). On average, this implies that the assumed

values for φp and φf are quite realistic.

27

Page 29: The Economic and Statistical Value of Forecast ... · 2For a special case, Aiolfi and Timmermann (2005) provide conditions under which the population MSFE of an equally-weighted

the benchmarks are not influenced by transaction costs. Interestingly, because these strategies imply that

for long periods corner solutions (in which 100% of the portfolio is destined to either stocks or bonds) are

obtained, the performance of Neely-Weller and Pesaran-Timmermann strategies are hardly affected. Finally,

the performance of the regime switching models is substantially worsened when tc is taken into account. For

instance, under λ = 1, the mean monthly return of the MSIH(4,0)-VAR(1) model goes from 1.12% to 0.88%;

correspondingly the Sharpe ratio declines from 0.211 to 0.162. In spite of this, the out-of-sample performance

of forecast-combination driven strategies remains overwhelmingly superior. Once more, assuming λ = 1, gives

a Sharpe ratio of 0.28 for the inverse-MSFE strategy vs. 0.16 for the zero-transaction cost portfolios and a

0.20 for Neeley-Weller and Pesaran-Timmermann’s strategies.

Table 11 approximately here

8. Conclusion

This paper has investigated two closely related issues. First, we have proposed to use multivariate regime

switching models to capture the presence of time-variation — structural shifts — in the predictability patterns

involving US asset returns and macroeconomic variables. Using a long monthly data set (1926-2004) we find

overwhelming evidence of regimes in the joint process for returns and macroeconomic factors. The good

performance of our four-state model at fitting the entire density of the data and its forecasting performance

stress that payoffs may exist in explicitly modeling the presence of regimes. Second, we have built a rather

strong case — both in statistical and economic terms — in favor of using forecast combinations in the presence

of regime switching, i.e. against treating data sets in which regimes are detected with the classical “horse

race” approach that aims at picking the best performing model(s) for forecasting purposes. On the opposite,

our results indicate that pooling schemes exist that both improve the prediction performance in (pseudo)

out-of-sample experiments, and that are of help to portfolio managers.

Interestingly, our results have also provided confirmation for a phenomenon often noted in the forecasting

literature: especially in the case of non-stationary data, to try to run regression and estimate optimal combi-

nation weights may be foolish. The underlying process is likely to be so complicated that the OLS estimates

are ridden of estimation error and as such hardly of any use. As our empirical analysis shows, the smart

choice becomes then to either adopt pooling schemes that refrain from estimation or — even better — adopt

mixed schemes that minimize the estimation needs and therefore maximize the chances of producing accurate

forecasts, such as Stock and Watson’s (2003) relative performance weights.

Several extensions of this paper appear to be rather natural. First of all, there is nothing compelling about

using multivariate regime switching models to study time-varying linkages between financial markets and the

macroeconomy. Other modeling approaches might prove useful, see e.g. Ravazzolo, Paap, van Dijk, and

Franses (2006). Second, for simplicity we have computed combination weights and assessed prediction accuracy

using an essentially univariate approach, in which — although some of the models under consideration are

multivariate in nature — the focus has been mainly directed to forecasting excess stock and bond returns, one

at the time. It would be interesting to consider whether adopting an explicit multivariate approach may change

the results found here. Third, there is ongoing theoretical research on the properties of alternative relative

performance schemes, which implies that our paper is likely to provide at best a lower bound to the gains that

may be delivered by “smartly” selected pooling designs. For instance, Yang (2004) studies (Bayesian) recursive

28

Page 30: The Economic and Statistical Value of Forecast ... · 2For a special case, Aiolfi and Timmermann (2005) provide conditions under which the population MSFE of an equally-weighted

forecast combination algorithms based on relative performance measures. The advantage of these methods is

ensure some degrees of minimal performance relative to the best (yet unknown) forecasting models. Finally,

our application in Section 7 to portfolio management merely scratches the surface of the relevant issues. For

instance, other natural loss functions may be involved by the practice of portfolio management for which recent

research has revealed that modeling non-stationarities may be crucial (see e.g., Guidolin and Timmermann

(2006b) for applications to value-at-risk computations).

References

[1] Aiolfi, M., and Timmermann, A. (2005). Persistence of Forecasting Performance and Combination Strate-

gies. Journal of Econometrics, forthcoming.

[2] Ang A., and Bekaert, G. (2002a). International Asset Allocation with Regime Shifts. Review of Financial

Studies 15, 1137-1187.

[3] Ang, A., and Bekaert, G. (2002b). Regime Switches in Interest Rates. Journal of Business and Economic

Statistics 20, 163-182.

[4] Bates, J., and Granger, C. (1969). The Combination of Forecasts. Operations Research Quarterly 20,

451-468.

[5] Bekaert, G., and Hodrick, R. (1992) Characterizing Predictable Components in Excess Returns on Equity

and Foreign Exchange Markets. Journal of Finance 47, 467-509.

[6] Berkowitz, J. (2001). Testing Density Forecasts with Applications to Risk Management. Journal of Busi-

ness and Economic Statistics 19, 465-474.

[7] Bossaerts, P., and Hillion, P. (1999). Implementing Statistical Criteria to Select Return Forecasting

Models: What Do We Learn? Review of Financial Studies 12, 405-428.

[8] Bunn, D. (1985). Statistical Efficiency in the Linear Combination of Forecasts. International Journal of

Forecasting 1, 151-163.

[9] Campbell, J. (1987). Stock Returns and the Term Structure. Journal of Financial Economics 18, 373-399.

[10] Chan, L., Karceski, J., and Lakonishok, J. (1998). The Risk and Return from Factors. Journal of Financial

and Quantitative Analysis 33, 159-188.

[11] Clemen, R. (1989). Combining Forecasts: A Review and Annotated Bibliography. International Journal

of Forecasting 5, 559-583.

[12] Clements M., and Hendry, D. (1998). Forecasting Economic Time Series. Cambridge University Press.

[13] Clements, M., and Krolzig, H. (1998). A Comparison of the Forecast Performance of Markov-Switching

and Threshold Autoregressive Models of US GNP. The Econometrics Journal 1, C47-C75.

[14] Clements, M., and Smith, J. (1997). The Performance of Alternative Forecasting Methods for SETAR

Models. International Journal of Forecasting 13, 463-475.

[15] Cutler, D., Poterba, J., and Summers, L. (1989). What Moves Stock Prices? Journal of Portfolio Man-

agement 15, 4-12.

29

Page 31: The Economic and Statistical Value of Forecast ... · 2For a special case, Aiolfi and Timmermann (2005) provide conditions under which the population MSFE of an equally-weighted

[16] Davies, R. (1977). Hypothesis Testing When a Nuisance Parameter Is Present Only Under the Alternative.

Biometrika 64, 247-254.

[17] Diebold, F. (1988). Serial Correlation and the Combination of Forecasts. Journal of Business and Eco-

nomic Statistics 6, 105-111.

[18] Diebold, F., Gunther, T., and Tay, A. (1998). Evaluating Density Forecasts. International Economic

Review 39, 863-883.

[19] Diebold, F., and Lopez, J. (1996). Forecast Evaluation and Combination. In G., Maddala, and C., Rao,

(Eds.), Handbook of Statistics 14 : Statistical Methods of Finance, Elsevier, 241-268.

[20] Diebold, F. , and Mariano, R. (1995). Comparing Predictive Accuracy. Journal of Business and Economic

Statistics 13, 253-263.

[21] Diebold, F., Rudebusch, G., and Sichel, D. (1993). Further Evidence on Business Cycle Duration Depen-

dence. In J., Stock and M., Watson (Eds.), Business Cycles, Indicators, and Forecasting, University of

Chicago Press and NBER, 255-280.

[22] Dunis, C., Laws, J., and Chauvin, S. (2001). The Use of Market Data and Model Combinations to

Improve Forecast Accuracy. In Dunis, C., A., Timmermann, and J., Moody (Eds.), Developments in

Forecast Combination and Portfolio Choice, John Wiley and Sons, Inc., 45-80.

[23] Elliott, G., and Timmermann, A. (2004). Optimal Forecast Combinations under General Loss Functions

and Forecast Error Distributions. Journal of Econometrics 122, 47-79.

[24] Fama, E., and French, K. (1988). Dividend Yields and Expected Stock Returns. Journal of Financial

Economics 19, 3-29.

[25] Fama, E., and French, K. (1989). Business Conditions and Expected Returns on Stocks and Bonds.

Journal of Financial Economics 25, 23-49.

[26] Fama, E., and French, K. (1993). Common Risk Factors in the Returns on Stocks and Bonds. Journal of

Financial Economics 33, 3-36.

[27] Fama, E., and Schwert, W. (1977). Asset Returns and Inflation. Journal of Financial Economics 5,

115-146.

[28] Fenton, V., and Gallant, R. (1996). Qualitative and Asymptotic Performance of SNP Density Estimation.

Journal of Econometrics 74, 77-118.

[29] Fugazza, C., Guidolin, M., and Nicodano, G. (2006). Investing for the Long-Run in European Real Estate.

Journal of Real Estate Finance and Economics, forthcoming.

[30] Garcia, R. (1998). Asymptotic Null Distribution of the Likelihood Ratio Test in Markov Switching Models.

International Economic Review 39, 763-788.

[31] Goyal, A., and Welch, I. (2003). Predicting the Equity Premium with Dividend Ratios. Management

Science 49, 639-654.

[32] Gray, S. (1996). Modeling the Conditional Distribution of Interest Rates as a Regime-Switching Process.

Journal of Financial Economics 42, 27-62.

30

Page 32: The Economic and Statistical Value of Forecast ... · 2For a special case, Aiolfi and Timmermann (2005) provide conditions under which the population MSFE of an equally-weighted

[33] Guidolin, M., and Ono, S. (2006). Are the Dynamic Linkages Between the Macroeconomy and Asset

Prices Time-Varying? Journal of Economics and Business 58, 480-518.

[34] Guidolin, M., and Timmermann, A. (2005a). Economic Implications of Bull and Bear Regimes in UK

Stock and Bond Returns. Economic Journal 115, 111-143.

[35] Guidolin, M., and Timmermann, A. (2005b). Optimal Forecast Combination Weights under Regime

Shifts: An Application to US Interest Rates. Mimeo, Federal Reserve Bank of St. Louis and UCSD.

[36] Guidolin, M., and Timmermann, A. (2006a). An Econometric Model of Nonlinear Dynamics in the Joint

Distribution of Stock and Bond Returns. Journal of Applied Econometrics 21, 1-22.

[37] Guidolin, M., and Timmermann, A. (2006b). Term Structure of Risk under Alternative Econometric

Specifications. Journal of Econometrics 131, 285-308.

[38] Guidolin, M., and Timmermann, A. (2006c). Asset Allocation under Regime Switching. Journal of Eco-

nomic Dynamics and Control, forthcoming.

[39] Hamilton, J. (1989). A New Approach to the Economic Analysis of Nonstationary Time Series and the

Business Cycle. Econometrica 57, 357-384.

[40] Hamilton, J., and Susmel, R. (1994). Autoregressive Conditional Heteroskedasticity and Changes in

Regime. Journal of Econometrics 64, 307-333.

[41] Hendry, D., and Clements, M. (2002). Pooling of Forecasts. Econometrics Journal 7, 1-31.

[42] James, C., Koreisha, S., and Partch, M. (1985). A VARMA Analysis of Casual Relations Among Stock

Returns, Real Output, and Nominal Interest Rates. Journal of Finance 40, 1375-1384.

[43] Jorion, P. (1985). Bayes-Stein Estimation for Portfolio Analysis. Journal of Financial and Quantitative

Analysis 21, 279-92.

[44] Krolzig, H.-M. (1997). Markov-Switching Vector Autoregressions. Berlin, Springer-Verlag.

[45] Lynch, A., and Balduzzi, P. (2000). Predictability and Transaction Costs: The Impact on Rebalancing

Rules and Behavior. Journal of Finance 55, 2285-2309.

[46] Neely, C., and Weller, P. (2000). Predictability in International Asset Returns: Reexamination. Journal

of Financial and Quantitative Analysis 35, 601-620.

[47] Palm, F., and Zellner, A. (1992). To Combine or Not to Combine? Issues of Combining Forecasts. Journal

of Forecasting 11, 687-701.

[48] Paye, B., and Timmermann, A. (2006). Instability of Return Prediction Models. Journal of Empirical

Finance, forthcoming.

[49] Pesaran, H., and Timmermann, A. (1995). Predictability of Stock Returns: Robustness and Economic

Significance. Journal of Finance 50, 1201-1228.

[50] Rapach, D., and Wohar, M. (2006). In-Sample vs. Out-of-Sample Tests of Stock Return Predictability in

the Context of Data Mining. Journal of Empirical Finance 13, 231-247.

31

Page 33: The Economic and Statistical Value of Forecast ... · 2For a special case, Aiolfi and Timmermann (2005) provide conditions under which the population MSFE of an equally-weighted

[51] Ravazzolo, F., Paap, R., van Dijk, D., and Franses, P., H. (2006). Bayesian Model Averaging in the

Presence of Structural Breaks. Mimeo, Erasmus University Rotterdam.

[52] Sola, M., and Driffill, J. (1994). Testing the Term Structure of Interest Rates Using a Stationary Vector

Autoregression with Regime Switching. Journal of Economic Dynamics and Control 18, 601-628.

[53] Stock, J., and Watson, M. (2001). A Comparison of Linear and Nonlinear Univariate Models for Fore-

casting Macroeconomic Time Series. In R., Engle, and H., White (Eds.), Cointegration, Causality, and

Forecasting: A Festschrift in Honour of Clive Granger, Oxford University Press, 1-44.

[54] Stock, J., and Watson, M. (1989). New Indexes of Coincident and Leading Economic Indicators. NBER

Macro-Economic Annual, 351-409.

[55] Stock, J., and Watson, M. (2003). Forecasting Output and Inflation: the Role of Asset Prices. Journal

of Economic Literature 41, 788-829.

[56] Timmermann, A. (2005). Forecast Combinations. Forthcoming in Handbook of Economic Forecasting.

Elsevier Press.

[57] Turner, C., Starz, R., and Nelson, C. (1989). A Markov Model of Heteroskedasticity, Risk, and Learning

in the Stock Market. Journal of Financial Economics 25, 3-22.

[58] De Miguel, V., Garlappi, L., and Uppal, R. (2006). 1/N. Mimeo, London Business School.

[59] Winkler, R. (1989). Combining Forecasts: A Philosophical Bias and Some Current Issues. International

Journal of Forecasting 5, 605-609.

[60] Winkler, R., and Makridakis, S. (1983). The Combination of Forecasts. Journal of the Royal Statistical

Society Series A 146, 150-157.

[61] Yang, Y. (2004). Combining Forecasts Procedures: Some Theoretical Results. Econometric Theory 20,

176-190.

32

Page 34: The Economic and Statistical Value of Forecast ... · 2For a special case, Aiolfi and Timmermann (2005) provide conditions under which the population MSFE of an equally-weighted

33

Figure 1

Quantile-Quantile Plots (vs. a Gaussian with identical mean and variance) for Transformed z-Scores from Three-State VAR(1) Switching Model

-4

-2

0

2

4

-4 -2 0 2 4

Excess stock returns z-scores

Nor

mal

Qua

ntile

-4

-2

0

2

4

-4 -2 0 2 4

Excess bond returns z-scores

Nor

mal

Qua

ntile

-10

-5

0

5

10

-15 -10 -5 0 5 10

Dividend yield z-scores

Nor

mal

Qua

ntile

-4

-2

0

2

4

6

-4 -2 0 2 4 6

T_bill z-scores

Nor

mal

Qua

ntile

-10

-5

0

5

10

-10 -5 0 5 10

Default spread z-scores

Nor

mal

Qua

ntile

-4

-2

0

2

4

-4 -2 0 2 4

CPI inflation z-scores

Nor

mal

Qua

ntile

-4

-2

0

2

4

-4 -2 0 2 4

Industrial production z-scores

Nor

mal

Qua

ntile

-4

-2

0

2

4

-4 -2 0 2 4

Monetary base z-scores

Nor

mal

Qua

ntile

Page 35: The Economic and Statistical Value of Forecast ... · 2For a special case, Aiolfi and Timmermann (2005) provide conditions under which the population MSFE of an equally-weighted

34

Figure 2

Quantile-Quantile Plots (vs. a Gaussian with identical mean and variance) for Transformed z-Scores from Four-State Model with Time-Invariant VAR(1) Matrix

-4

-2

0

2

4

-4 -2 0 2 4

Excess stock returns z-scores

Nor

mal

Qua

ntile

-4

-2

0

2

4

-4 -2 0 2 4

Excess bond returns z-scores

Nor

mal

Qua

ntile

-4

-2

0

2

4

-4 -2 0 2 4

Dividend yield z-score

Nor

mal

Qua

ntile

-4

-2

0

2

4

-4 -2 0 2 4

T-bill z-scores

Nor

mal

Qua

ntile

-4

-2

0

2

4

-4 -2 0 2 4

Default spread z-scores

Nor

mal

Qua

ntile

-4

-2

0

2

4

-4 -2 0 2 4

Inflation z-scores

Nor

mal

Qua

ntile

-4

-2

0

2

4

-4 -2 0 2 4

Industrial production z-scores

Nor

mal

Qua

ntile

-4

-2

0

2

4

-4 -2 0 2 4

Monetary base z-scores

Nor

mal

Qua

ntile

Page 36: The Economic and Statistical Value of Forecast ... · 2For a special case, Aiolfi and Timmermann (2005) provide conditions under which the population MSFE of an equally-weighted

35

Figure 3

Smoothed Probabilities from Four-State Model

0.0

0.2

0.4

0.6

0.8

1.0

1930 1940 1950 1960 1970 1980 1990 2000

Regime 1

0.0

0.2

0.4

0.6

0.8

1.0

1930 1940 1950 1960 1970 1980 1990 2000

Regime 2

0.0

0.2

0.4

0.6

0.8

1.0

1930 1940 1950 1960 1970 1980 1990 2000

Regime 3

0.0

0.2

0.4

0.6

0.8

1.0

1930 1940 1950 1960 1970 1980 1990 2000

Regime 4

Page 37: The Economic and Statistical Value of Forecast ... · 2For a special case, Aiolfi and Timmermann (2005) provide conditions under which the population MSFE of an equally-weighted

36

Figure 4

Comparing Fitted and Realized Values: Four-State Model with Time-Invariant VAR(1) Matrix

-.4

-.2

.0

.2

.4

-.04

-.02

.00

.02

.04

.06

.08

1930 1940 1950 1960 1970 1980 1990 2000

Excess stock returns Risk premium MSIH(4,0)-VAR Risk premium VAR(1)

Correlation actual - MSIH(4,1)-VAR: 0.23Correlation actual - VAR(1): 0.17

-.08

-.04

.00

.04

.08

.12

-.02

-.01

.00

.01

.02

1930 1940 1950 1960 1970 1980 1990 2000

Excess bond returns Excess bond MSIH(4,0)-VAR Excess bond returns VAR(1)

Correlation actual - MSIH(4,1)-VAR: 0.14Correlation actual - VAR(1): 0.12

.00

.04

.08

.12

.16

.00

.04

.08

.12

.16

1930 1940 1950 1960 1970 1980 1990 2000

Dividend yieldDividend yield MSIH(4,0)-VAR Dividend yield VAR(1)

Correlation actual - MSIH(4,1)-VAR: 0.98Correlation actual - VAR(1): 0.98

-.04

-.02

.00

.02

.04

-.02

-.01

.00

.01

.02

1930 1940 1950 1960 1970 1980 1990 2000

T-bill yieldT-bill yield MSIH(4,0)-VART-bill yield VAR(1)

Correlation actual - MSIH(4,1)-VAR: 0.61Correlation actual - VAR(1): 0.50

.000

.001

.002

.003

.004

.005

.000

.001

.002

.003

.004

.005

1930 1940 1950 1960 1970 1980 1990 2000

Default spreadDefault spread MSIH(4,0)-VARDefault spread VAR(1)

Correlation actual - MSIH(4,1)-VAR: 0.98Correlation actual - VAR(1): 0.96

-.04

-.02

.00

.02

.04

-.012

-.008

-.004

.000

.004

.008

.012

.016

1930 1940 1950 1960 1970 1980 1990 2000

CPI inf lationCPI inf lation MSIH(4,0)-VAR CPI inf lation VAR(1)

Correlation actual - MSIH(4,1)-VAR: 0.65Correlation actual - VAR(1): 0.53

-.2

-.1

.0

.1

.2

-.04

.00

.04

.08

.12

1930 1940 1950 1960 1970 1980 1990 2000

IP grow th rateIP grow th rate MSIH(4,0)-VAR IP grow th rate VAR(1)

Correlation actual - MSIH(4,1)-VAR: 0.52Correlation actual - VAR(1): 0.53

-.1

.0

.1

.2

.3

-.04

-.02

.00

.02

.04

.06

.08

1930 1940 1950 1960 1970 1980 1990 2000

Mon. base grow thMon. base gr. MSIH(4,0)-VAR Mon. base grow th VAR(1)

Correlation actual - MSIH(4,1)-VAR: 0.30Correlation actual - VAR(1): 0.22

Page 38: The Economic and Statistical Value of Forecast ... · 2For a special case, Aiolfi and Timmermann (2005) provide conditions under which the population MSFE of an equally-weighted

37

Figure 5

Recursive Combination Weights: One-Month Horizon

Panel A: OLS weights

Excess Stock Returns: MSIH(4,0)-VAR(1) + PT Rbarsqaure-selection

-2

0

2

4

6

8

10

12

14

1986 1988 1990 1992 1994 1996 1998 2000 2002 2004

Intercept MSIH(4.0)-VAR(1) weight PT Rbarsqaure-selection weight

Excess Bond Returns: MSIH(4,0)-VAR(1) + PT Rbarsqaure-selection

-0.5

-0.3

-0.1

0.1

0.3

0.5

0.7

0.9

1986 1988 1990 1992 1994 1996 1998 2000 2002 2004

Intercept MSIH(4.0)-VAR(1) weight PT Rbarsquare-selection weight

Panel B: Inverse-MSFE weights

Excess Stock Returns

0.1

0.13

0.16

0.19

0.22

0.25

1986 1988 1990 1992 1994 1996 1998 2000 2002 2004MSIH(4,0)-VAR(1) MSIAH(3,1)VAR(1) Neely-Weller's benchmarkPT-Adj. Rsquare selection PT BIC-selectionAR BIC-selection

Excess Bond Returns

0.1

0.12

0.14

0.16

0.18

1986 1988 1990 1992 1994 1996 1998 2000 2002 2004

MSIH(4,0)-VAR(1) MSIAH(3,1)VAR(1) Neely-Weller's benchmarkPT - Adj. Rsquare selection PT BIC-selectionAR BIC-selection

Page 39: The Economic and Statistical Value of Forecast ... · 2For a special case, Aiolfi and Timmermann (2005) provide conditions under which the population MSFE of an equally-weighted

38

Figure 6

Implied Monthly Predicted Sharpe Ratios from Different Models

One-Month Ahead Sharpe Ratios on Stocks

-0.6

-0.5

-0.4

-0.3

-0.2

-0.1

0

0.1

0.2

0.3

0.4

0.5

1985 1987 1989 1991 1993 1995 1997 1999 2001 2003 2005

MSIH(4,0)-VAR(1) MSIAH(3,1)VAR(1) Neely-Weller's benchmarkPT - Adj. R-square selection PT - BIC selectionAR models - BIC selection

One-Month Ahead Sharpe Ratios on Bonds

-0.6

-0.5

-0.4

-0.3

-0.2

-0.1

0

0.1

0.2

0.3

0.4

0.5

1985 1987 1989 1991 1993 1995 1997 1999 2001 2003 2005

MSIH(4,0)-VAR(1) MSIAH(3,1)VAR(1) Neely-Weller's benchmarkPT - Adj. R-square selection PT - BIC selectionAR models - BIC selection

Page 40: The Economic and Statistical Value of Forecast ... · 2For a special case, Aiolfi and Timmermann (2005) provide conditions under which the population MSFE of an equally-weighted

39

Table 1

Summary Statistics for Excess Stock and Bond Returns vs. Prediction Variables The table reports a few summary statistics for monthly CRSP excess stock and (long-term) government bond return series, and macroeconomic variables employed as predictors of excess asset returns. Excess returns are calculated by difference with 30-day T-bill yields. The sample period is 1926:12 – 2004:12. In the case of equities, the CRSP universe spans stocks listed on the NYSE, the NASDAQ, and the AMEX. Data on bond returns refer to the CRSP 10-Year Treasury benchmark. All returns are expressed in monthly percentage terms. LB(j) denotes the j-th order Ljung-Box statistic.

Series Mean Median St. Dev. Skewness Kurtosis Jarque-Bera

LB(4) LB(4)- squares

Excess Asset Returns (Risk Premia) Value-weighted excess stock returns 0.6482 0.9900 5.4946 0.2133 10.6124 2269** 21.716** 166.87**

Excess bond returns (term premium) 0.1447 0.1400 1.8808 0.2447 5.5932 271.9** 5.1774 176.31**

Prediction Variables 12-month cumulated dividend yield 3.8132 3.6340 1.4987 0.9542 5.8183 452.3** 3334** 2829**

Real 1-month T-bill yield 0.0540 0.0700 0.5114 -1.9764 21.0381 13313** 542.13** 79.833**

Default spread 0.0943 0.0730 0.0600 2.4203 11.3805 3657** 3284** 2683** CPI inflation rate 0.2498 0.2659 0.5279 1.1840 16.7930 7647** 596.9** 82.741**

Industrial production growth rate 0.2101 0.2270 2.0208 0.7663 13.2813 4219** 268.7** 372.7**

Real adj. monetary base growth rate 0.0381 0.1540 2.3031 1.7034 30.7269 30468** 34.722** 79.514**

* denotes 5% significance, ** significance at 1%.

Page 41: The Economic and Statistical Value of Forecast ... · 2For a special case, Aiolfi and Timmermann (2005) provide conditions under which the population MSFE of an equally-weighted

40

Table 2 – part a

Estimates of a Four-State Switching Model with Time-Invariant VAR(1) Matrix Panel A – Single State VAR(1) Model Stock Bond Div. yield T-bill Default Inflation Growth Money 1. Intercept -0.3878

(0.6076) -0.0376

(0.2092) 0.1022***

(0.0331) -0.0425

(0.0522) 0.0001

(0.0013) 0.0605

(0.0518) 0.1356

(0.1910) 0.5340*

(0.3037) 2. VAR(1) Matrix Stock excess returns 0.1046***

(0.0333) 0.1108

(0.0960) 0.2653*

(0.1430) -0.2514

(0.7425) 0.5962

(3.5335) -0.3426

(0.7735) 0.0653

(0.0974) 0.1257**

(0.0586)

Bond excess returns -0.0260**

(0.0115) 0.0626*

(0.0331) -0.0020

(0.0492) 0.0522

(0.2557) 2.2183*

(1.2168) -0.0485

(0.2664) 0.0115

(0.0335) -0.0104

(0.0202)

Dividend yield -0.0060***

(0.0018) -0.0011

(0.0052) 0.9835***

(0.0078) -0.0403

(0.0405) -0.3175*

(0.1925) -0.0285

(0.0421) -0.0029

(0.0053) -0.0052*

(0.0032)

T-bill real yield -0.0011

(0.0029) 0.0013

(0.0083) -0.0249*

(0.0123) 0.5768***

(0.0638) 1.2793***

(0.3037) 0.1068*

(0.0605) -0.0518***

(0.0084) -0.0027

(0.0050)

Default spread -0.0009***

(0.0001) 0.0007***

(0.0002) 0.0007**

(0.0003) 0.0021

(0.0016) 0.9628***

(0.0077) 0.0002

(0.0017) -0.0005**

(0.0002) 0.0000

(0.0001)

Inflation 0.0013

(0.0028) -0.0087

(0.0082) 0.0232*

(0.0122) 0.3867***

(0.0633) -1.2891***

(0.3011) 0.8632***

(0.0659) 0.0526***

(0.0083) 0.0026

(0.0050)

IP real growth 0.0760***

(0.0105) -0.0372

(0.0302) -0.0854*

(0.0449) -0.9677***

(0.2334) 3.3244***

(1.1108) -0.9506***

(0.2432) 0.4008***

(0.0306) 0.0396**

(0.0184)

Money real growth -0.0025

(0.0188) 0.0219

(0.0543) -0.2341***

(0.0809) -1.3385***

(0.4201) 10.338***

(1.9992) -2.3610***

(0.4377) -0.0316

(0.0551) -0.1880***

(0.0332) 3. Correlations/Volatilities Stock excess returns 0.0542*** Bond excess returns 0.1377** 0.0187*** Dividend yield -0.8830*** -0.1368** 0.0030** T-bill real yield -0.0333 0.0561 0.0404 0.0047*** Default spread -0.2596*** 0.0712 0.3324*** 0.0371 0.0001* Inflation 0.0221 -0.0582 -0.0313 -0.9909*** -0.0298 0.0046*** IP real growth 0.1128** -0.0104 -0.1012** 0.4401*** -0.1744** -0.4434*** 0.0170*** Money real growth -0.0034 0.0594* -0.0027 0.2329** 0.0677* -0.2332** 0.0259 0.0031** Panel B – Four State Model Stock Bond Div. yield T-bill Default Inflation Growth Money 1. Intercept Bull-rebound 0.8945**

(0.4420) 0.2530***

(0.1031) 0.0382*

(0.0185) -0.2271***

(0.0781) 0.0022

(0.0046) 0.2321**

(0.1181) 0.1633*

(0.1090) -0.9862**

(0.3959)

Stable-growth 0.6454**

(0.2707) 0.2955***

(0.0957) 0.0169

(0.0173) -0.0008

(0.0416) 0.0009

(0.0019) 0.0225

(0.0418) 0.2330*

(0.1336) -0.0108

(0.2049)

Expansion-peak 0.5782***

(0.0874) 1.0141

(0.5851) 0.0408*

(0.0206) 0.0015

(0.0776) 0.0032

(0.0019) 0.0399

(0.0777) -0.0207

(0.2502) 0.0478*

(0.0261)

Bear-recession -1.9266***

(0.4820) -0.2598***

(0.0570) 0.1301***

(0.0095) 0.1575***

(0.0191) 0.0089*

(0.0052) -0.1582**

(0.0793) 0.4802***

(0.1830) 0.4399***

(0.0836) 2. VAR(1) Matrix Stock excess returns -0.0009

(0.0287) 0.1057

(0.0719) 0.0870

(0.1220) -1.3822***

(0.4207) 4.7186***

(0.0805) -2.1260***

(0.7250) -0.0433**

(0.0207) 0.0821*

(0.0450)

Bond excess returns -0.0341*

(0.0173) 0.0167

(0.0324) -0.0401*

(0.0242) -0.9605***

(0.2579) 2.9146***

(1.0360) -1.0749***

(0.3645) -0.0079

(0.0204) 0.0159

(0.0137)

Dividend yield -0.0003

(0.0011) -0.0034

(0.0025) 0.9940***

(0.0044) 0.0343

(0.0298) -0.3475***

(0.0753) 0.0619*

(0.0302) 0.0032

(0.0038) -0.0044*

(0.0024)

T-bill real yield 0.0004

(0.0025) 0.0045*

(0.0024) 0.0011

(0.0102) 0.4529***

(0.0651) 0.3641**

(0.1875) 0.0811

(0.0669) -0.0187

(0.0092) -0.0025

(0.0055)

Default spread -0.0003*

(0.0002) 0.0003*

(0.0002) 0.0001

(0.0004) 0.0034*

(0.0022) 0.9640***

(0.0043) 0.0031*

(0.0020) 0.0000

(0.0001) -0.0000

(0.0001)

Inflation -0.0006

(0.0024) -0.0092

(0.0054) -0.0008

(0.0102) 0.4872***

(0.0655) -0.3839**

(0.1674) 0.8599***

(0.0672) 0.0190

(0.0142) 0.0028*

(0.0016)

IP real growth 0.0235*

(0.0120) -0.0089

(0.0183) -0.0385

(0.0362) -0.8952

(0.2186) 0.9575**

(0.4569) -0.9208***

(0.2189) 0.3861***

(0.0326) 0.0185

(0.0199)

Money real growth 0.0143

(0.0118) 0.0233

(0.0253) 0.0394

(0.0491) -1.4162***

(0.2968) 6.8721***

(1.9272) -2.3985***

(0.3015) -0.0294

(0.0464) -0.1786***

(0.0324) * denotes 10% significance, ** significance at 5%, *** significance at 1%.

Page 42: The Economic and Statistical Value of Forecast ... · 2For a special case, Aiolfi and Timmermann (2005) provide conditions under which the population MSFE of an equally-weighted

41

Table 2 – part b

Estimates of a Four-State Switching Model with Time-Invariant VAR(1) Matrix Panel B – Four State Model Stock Bond Div. yield T-bill Default Inflation Growth Money 3. Correlations/Volatilities Regime 1 (Bull-rebound): Stock excess returns 0.0385*** Bond excess returns 0.1494** 0.0068*** Dividend yield -0.8806*** -0.1687*** 0.0021*** T-bill real yield 0.0438 0.0522 -0.0286 0.0071*** Default spread -0.1434** -0.0127 0.1049** -0.0593 3.9e-05* Inflation -0.0429 -0.0522 0.0270 -0.9990*** 0.0589 0.0071*** IP real growth -0.0918** 0.0388 0.0678* 0.5054*** -0.1040* -0.5051*** 0.0256*** Money real growth 0.1742** 0.0220 -0.1864** 0.3201*** -0.0891* -0.3211*** 0.1508** 0.0351***

Regime 2 (Stable-growth): Stock excess returns 0.0365*** Bond excess returns 0.1433** 0.0180*** Dividend yield -0.8999*** -0.1807*** 0.0011*** T-bill real yield 0.0784 0.0593 -0.0721 0.0023*** Default spread -0.0271 -0.0144 0.0204 0.0989* 3.98e-05* Inflation -0.0917* -0.0554 0.0884* -0.9857*** -0.1034** 0.0023*** IP real growth 0.0953* -0.0492 -0.1158** 0.5290*** -0.0144 -0.5406*** 0.0083*** Money real growth 0.0539 0.0009 -0.0449 0.3928*** 0.0472 -0.3953*** 0.2029** 0.0108***

Regime 3 (Expansion-peak): Stock excess returns 0.0606*** Bond excess returns 0.1965*** 0.0277*** Dividend yield -0.9691*** -0.2080*** 0.0026*** T-bill real yield 0.0063 0.0896 -0.0427 0.0043*** Default spread 0.0773* 0.3746*** -0.0809** 0.0494 0.0001** Inflation -0.0665 -0.1141* 0.1161** -0.9535*** -0.0045 0.0041*** IP real growth 0.1126** -0.0253 -0.1372** 0.5943*** -0.0540 -0.5789*** 0.0125*** Money real growth -0.0977** 0.0234 0.0702* 0.3791*** -0.0690 -0.4093*** 0.2576** 0.0198***

Regime 4 (Bear-recession): Stock excess returns 0.1196*** Bond excess returns 0.1040* 0.0197*** Dividend yield -0.9225*** -0.1873*** 0.0082*** T-bill real yield -0.1698** 0.1577** 0.1597** 0.0067*** Default spread -0.4836*** -0.1435** 0.4864*** 0.0744 0.0003*** Inflation 0.1727** -0.1605** -0.1631** -0.9981*** -0.0797* 0.0067*** IP real growth 0.3401*** 0.0683 -0.2432*** 0.1544** -0.4347*** -0.1601** 0.0317*** Money real growth -0.0456 0.2287*** 0.0456 0.0262 0.1042* -0.0180 -0.1768** 0.0769***

4. Transition probabilities Bull-rebound Stable-peak Expansion Bear-recession Bull-rebound 0.8939***

(0.1560) 0.0237*

(0.0128) 0.0207

(0.0249) 0.0616

Stable-growth 0.0100

(0.0349) 0.9311***

(0.2305) 0.0523**

(0.2410) 0.0066

Expansion-peak 0.0230

(0.2340) 0.1517***

(0.0555) 0.8030***

(0.1159) 0.0381

Bear-recession 0.1683***

(0.0200) 0.0340*

(0.0183) 0.0381**

(0.0160) 0.7596

* denotes 10% significance, ** significance at 5%, *** significance at 1%.

Page 43: The Economic and Statistical Value of Forecast ... · 2For a special case, Aiolfi and Timmermann (2005) provide conditions under which the population MSFE of an equally-weighted

42

Table 3

Out-of-Sample, Recursive Predictive Performance The table reports the root-mean-square forecast error, the predictive bias, and the forecast error standard deviation for the seven models described in the main text. The (pseudo) out-of sample period is 1985:01 – 2004:12. The models are recursively estimated on expanding windows 1926:12 – 1985:01, 1926:12 – 1985:02, up to 1926:12 – 2004:12 - h.

Stock Bond Div. yield T-bill Default Inflation Growth Money Four-state MSIH(4,0)-VAR(1) Root-MSFE 3.773 1.285 1.443 0.880 0.152 0.277 1.385 2.421 Bias 0.369 0.218 -1.428 -0.128 0.113 -0.224 0.359 0.561 St. dev. 3.755 1.266 0.208 0.871 0.102 0.163 1.338 2.355 Three-state MSIAH(4,1) Root-MSFE 4.310 1.287 3.120 0.888 0.115 0.927 1.422 5.570 Bias 0.883 0.287 2.533 -0.145 0.071 -0.167 0.318 0.266 St. dev. 4.219 1.255 1.821 0.876 0.090 0.912 1.386 5.564 Single-state VAR(1) Root-MSFE 5.256 1.257 1.580 0.900 0.115 0.253 2.442 2.964 Bias 0.631 0.227 -1.333 -0.124 0.071 -0.080 0.408 0.282 St. dev. 5.218 1.236 0.848 0.891 0.090 0.240 2.408 2.951 Neely and Weller’s benchmark Root-MSFE 6.138 1.238 3.163 0.891 0.115 0.923 3.438 5.881 Bias 0.036 0.146 2.578 -0.144 0.071 -0.169 0.456 0.599 St. dev. 6.138 1.229 1.832 0.879 0.091 0.907 3.408 5.850 Pesaran and Timmermann – Adj. R2 selection Root-MSFE 4.485 2.152 2.114 0.274 0.006 0.266 0.893 4.311 Bias 0.031 0.228 -0.008 0.000 -0.000 -0.002 0.009 -0.016 St. dev. 4.485 2.140 2.114 0.274 0.006 0.266 0.893 4.311 Pesaran and Timmermann – BIC selection Root-MSFE 4.561 2.168 2.598 0.260 0.023 0.250 0.709 3.007 Bias -0.859 0.270 -2.451 0.0338 -0.006 0.085 -0.104 0.551 St. dev. 4.480 2.151 0.861 0.258 0.022 0.235 0.702 2.956 AR(p) univariate models (BIC Selection) Root-MSFE 4.485 2.152 2.192 0.309 0.022 0.377 0.793 3.011 Bias -0.130 0.209 -2.017 -0.174 0.004 0.295 -0.370 -0.482

On

e-m

onth

hor

izon

St. dev. 4.483 2.142 0.858 0.256 0.022 0.235 0.702 2.972 Four-state MSIH(4,0)-VAR(1) Root-MSFE 4.560 2.129 0.478 0.272 0.017 0.252 0.739 3.001 Bias 0.429 0.266 -0.297 0.058 -0.006 -0.063 -0.109 0.003 St. dev. 4.540 2.112 0.374 0.266 0.016 0.243 0.731 3.001 Three-state MSIAH(4,1) Root-MSFE 5.076 2.385 0.830 0.300 0.022 0.291 0.935 3.347 Bias 0.273 0.437 -0.628 -0.092 -0.010 0.081 -0.375 -0.318 St. dev. 5.068 2.345 0.543 0.286 0.020 0.279 0.856 3.332 Single-state VAR(1) Root-MSFE 5.543 2.882 0.882 1.033 0.027 1.109 4.272 4.771 Bias -1.076 1.308 0.478 -0.816 0.001 0.894 -3.491 -2.762 St. dev. 5.438 2.568 0.741 0.633 0.027 0.656 2.463 3.890 Random Walk benchmarks Root-MSFE 5.269 2.452 0.400 0.329 0.021 0.320 0.973 4.788 Bias 0.007 0.181 -0.108 -0.025 -0.003 -0.002 0.017 -0.069 St. dev. 5.269 2.445 0.385 0.328 0.021 0.320 0.973 4.788 Pesaran and Timmermann – Adj. R2 selection Root-MSFE 4.695 2.184 2.395 0.278 0.023 0.264 0.763 3.102 Bias -0.411 0.214 -2.153 0.068 -0.010 -0.024 -0.227 -0.377 St. dev. 4.677 2.173 1.047 0.269 0.021 0.263 0.728 3.078 Pesaran and Timmermann – BIC selection Root-MSFE 4.639 2.149 2.406 0.272 0.023 0.262 0.735 3.070 Bias -0.352 0.192 -2.192 0.062 -0.010 -0.018 -0.157 -0.442 St. dev. 4.627 2.140 0.992 0.265 0.021 0.262 0.719 3.038 AR(p) univariate models (BIC Selection) Root-MSFE 4.650 2.164 2.163 0.267 0.021 0.258 0.728 3.138 Bias 0.013 0.182 -1.961 0.034 -0.004 0.080 -0.040 -0.437

On

e-ye

ar h

oriz

on

St. dev. 4.650 2.156 0.912 0.264 0.021 0.245 0.727 3.107

Page 44: The Economic and Statistical Value of Forecast ... · 2For a special case, Aiolfi and Timmermann (2005) provide conditions under which the population MSFE of an equally-weighted

43

Table 4

Predictive Accuracy Tests – 1-Month Horizon The table reports Diebold-Mariano test statistics for pairwise comparisons of the MSE produced by different forecasting models. The test is applied to forecast errors from recursive, 1-step ahead forecasts of excess stock and bond returns using a variety of models. Numbers in bold highlight pairs of models for which one forecast function significantly outperforms the other at a 5% level using Diebold and Mariano (1995) asymptotic normal distribution. Numbers in parenthesis are p-values obtained from an application of the block bootstrap to MSE differentials (based on 50,000 independent block trials). Statistics illustrate the comparative forecasting performance of the model in the row vs. the model in the column: a negative (positive) value indicates that the row model out- (under-) performs the column model. Below the main diagonal, we present results for excess stock returns; above the main diagonal, results are for excess bond returns. PT – Adj. R2 selection PT – BIC

selection AR - BIC selection

Random walk benchmark VAR(1) MSIAH(3,1) MSIH(4,0)-

VAR(1)

`Excess Bond Returns

PT – Adj. R2 selec`tion -2.719 (0.143)

0.126 (0.806)

3.505 (0.006)

1.028 (0.278)

1.063 (0.327)

7.273 (0.003)

PT – BIC selection 0.586 (0.471)

2.175 (0.074)

2.719 (0.042)

1.288 (0.251)

1.172 (0.105)

7.521 (0.002)

AR - BIC selection 0.022 (0.098)

-1.579 (0.468)

2.126 (0.181)

1.013 (0.346)

1.030 (0.043)

7.222 (0.007)

Random walk 3.405 (0.034)

1.299 (0.270)

0.022 (0.819)

-0.028 (0.845)

-1.967 (0.144)

-0.294 (0.674)

VAR(1) 0.786 (0.410)

0.092 (0.167)

0.712 (0.404)

0.922 (0.317)

-6.442 (0.038)

-0.956 (0.510)

MSIAH(3,1) -2.297 (0.048)

-2.697 (0.123)

-2.235 (0.080)

-0.998 (0.275)

-2.002 (0.091)

0.125 (0.923)

MSIH(4,0)-VAR(1)

Exc

ess

Stoc

k R

etu

rns

-6.326 (0.000)

-5.390 (0.021)

-5.519 (0.046)

-1.817 (0.157)

-6.275 (0.031)

-2.255 (0.065)

Page 45: The Economic and Statistical Value of Forecast ... · 2For a special case, Aiolfi and Timmermann (2005) provide conditions under which the population MSFE of an equally-weighted

44

Table 5

Predictive Accuracy Tests – 12-Month Horizon The table reports Diebold-Mariano test statistics for pairwise comparisons of the MSE produced by different forecasting models. The test is applied to forecast errors from recursive, 12-step ahead forecasts of excess stock and bond returns using a variety of models. Numbers in bold highlight pairs of models for which one forecast function significantly outperforms the other at a 5% level using Diebold and Mariano (1995) asymptotic normal distribution. Numbers in parenthesis are p-values obtained from an application of the block bootstrap to MSE differentials (based on 50,000 independent block trials). Statistics illustrate the comparative forecasting performance of the model in the row vs. the model in the column: a negative (positive) value indicates that the row model out- (under-) performs the column model. Below the main diagonal, we present results for excess stock returns; above the main diagonal, results are for excess bond returns. PT – Adj. R2 selection PT – BIC

selection AR - BIC selection

Random walk benchmark VAR(1) MSIAH(3,1) MSIH(4,0)-

VAR(1)

`Excess Bond Returns

PT – Adj. R2 selec`tion 2.640 (0.049)

1.657 (0.103)

1.671 (0.099)

-3.458 (0.009)

-1.323 (0.283)

1.278 (0.203)

PT – BIC selection -1.237 (0.274)

0.330 (0.794)

1.352 (0.169)

-3.565 (0.034)

-2.007 (0.060)

1.818 (0.084)

AR - BIC selection -1.709 (0.104)

1.472 (0.241)

1.603 (0.103)

-3.485 (0.010)

-1.822 (0.084)

1.628 (0.150)

Random walk 1.602 (0.088)

1.450 (0.184)

2.968 (0.030)

-1.825 (0.085)

-4.854 (0.005)

3.850 (0.009)

VAR(1) 1.770 (0.083)

1.690 (0.140)

1.147 (0.359)

1.157 (0.320)

4.367 (0.017)

3.647 (0.024)

MSIAH(3,1) 1.666 (0.134)

1.610 (0.114)

1.553 (0.145)

1.509 (0.104)

0.357 (0.749)

3.586 (0.011)

MSIH(4,0)-VAR(1)

Exc

ess

Stoc

k R

etu

rns

-2.056 (0.059)

-1.979 (0.063)

-2.072 (0.047)

-0.836 (0.291)

-2.353 (0.050)

-0.637 (0.536)

Page 46: The Economic and Statistical Value of Forecast ... · 2For a special case, Aiolfi and Timmermann (2005) provide conditions under which the population MSFE of an equally-weighted

45

Table 6

Recursive Predictive Performance of Forecast Combinations – One Month Horizon The table reports the root-mean-square forecast error, the predictive bias, and the forecast error standard deviation for a number of alternative forecast combination schemes. The first panel concerns combinations in pairs, when a baseline forecast (i.e. forecast 1) is held fixed at the forecast function listed in the header. Pairwise weights are estimated by ordinary least squares. The second panel groups a few benchmarks: (i) when all forecast functions are recursively combined (with least-squares weights); (ii) when weights are fixed to be identical across all forecast functions; Stock and Watson’s (2003) scheme in which weights are inversely proportional to the recursive sample MSFE of each forecast function. The (pseudo) out-of sample period is 1985:01 – 2004:11.

Excess Stock Returns Excess Bond Returns Root-MSFE Bias Std. Dev. Root-MSFE Bias Std. Dev.

Combination with:

Panel A1: Baseline forecast: MSIH(4,0)-VAR(1)

MSIAH(3,1) 6.141 -0.006 6.141 2.812 0.005 2.812 VAR(1) 6.133 0.028 6.133 2.808 0.010 2.808

Random walk bench. 6.130 -0.015 6.130 2.810 -0.014 2.810

PT – Adj. R2 selection

6.136 -0.013 6.136 2.807 0.010 2.806

PT – BIC selection

6.134 -0.007 6.134 2.807 0.009 2.807

AR - BIC selection

6.134 -0.031 6.134 2.811 -0.002 2.811

Panel A2: Other benchmarks Baseline

MSIH(4,0)-VAR(1) 3.773 0.369 3.755 1.285 0.218 1.266 All forecast functions 5.026 0.345 5.014 0.439 -0.014 0.439

Equally weighted 2.169 -0.303 2.148 1.047 0.106 1.042 Inv. MSFE-

weighted 1.567 -0.301 1.538 0.689 0.057 0.687

Panel B1: Baseline forecast: Pesaran and Timmermann – Adj. R2 selection MSIAH(3,1) 6.096 -0.121 6.094 2.822 -0.004 2.822

VAR(1) 4.653 0.375 4.638 2.212 -0.165 2.206 Neely-Weller

bench. 4.593 0.552 4.560 2.172 0.194 2.164

MSIH(4,0)-VAR(1)

6.136 -0.013 6.136 2.806 0.010 2.806

PT – BIC selection

4.646 0.168 4.643 2.129 -0.100 2.126

AR - BIC selection

4.588 0.153 4.586 2.173 0.089 2.171

Panel B2: Other benchmarks Baseline PT –

Adj. R2 selection 4.485 0.031 4.627 2.152 0.228 2.140

All forecast functions 5.026 0.345 5.014 0.439 -0.014 0.439

Equally weighted 2.169 -0.303 2.148 1.047 0.106 1.042 Inv. MSFE-

weighted 1.567 -0.301 1.538 0.689 0.057 0.687

Page 47: The Economic and Statistical Value of Forecast ... · 2For a special case, Aiolfi and Timmermann (2005) provide conditions under which the population MSFE of an equally-weighted

46

Table 7

Recursive Predictive Performance of Forecast Combinations – 12-Month Horizon The table reports the root-mean-square forecast error, the predictive bias, and the forecast error standard deviation for a number of alternative forecast combination schemes. The first panel concerns combinations in pairs, when a baseline forecast (i.e. forecast 1) is held fixed at the forecast function listed in the header. Pairwise weights are estimated by ordinary least squares. The second panel groups a few benchmarks: when all forecast functions are recursively combined (with least-squares weights); when weights are fixed to be identical across all forecast functions; Stock and Watson’s (2003) scheme in which weights are inversely proportional to the recursive sample MSFE of each forecast function. The (pseudo) out-of sample period is 1985:01 – 2003:12.

Excess Stock Returns Excess Bond Returns Root-MSFE Bias Std. Dev. Root-MSFE Bias Std. Dev.

Combination with:

Panel A: Baseline forecast: MSIH(4,0)-VAR(1)

MSIAH(3,1) 4.665 -0.041 4.665 2.262 0.031 2.262 VAR(1) 4.663 -0.082 4.662 2.240 -0.014 2.240

Neely-Weller bench. 4.561 -0.044 4.561 2.276 0.386 2.243

PT – Adj. R2 selection

4.550 0.146 4.548 2.269 0.045 2.268

PT – BIC selection

4.653 0.033 4.653 2.262 0.015 2.262

AR - BIC selection

4.579 -0.037 4.579 2.265 0.357 2.237

Panel A: Other benchmarks Baseline

MSIH(4,0)-VAR(1) 4.560 2.129 4.540 2.129 0.266 2.112 All forecast functions 5.234 -1.340 5.060 3.277 -0.018 3.277

Equally weighted 4.485 -0.124 4.483 2.093 0.400 2.054 Inv. MSFE-

weighted 4.100 -0.146 4.097 1.924 0.358 1.890

Panel B: Baseline forecast: Pesaran and Timmermann – BIC selection MSIAH(3,1) 4.690 -0.087 4.689 2.232 0.031 2.232

VAR(1) 4.763 -0.048 4.763 2.238 0.056 2.238 Neely-Weller

bench. 4.692 -0.022 4.692 2.312 0.358 2.284

MSIH(4,0)-VAR(1)

4.691 0.017 4.691 2.228 0.002 2.228

PT – BIC selection

4.653 0.033 4.653 2.262 0.015 2.262

AR - BIC selection

4.698 -0.002 4.698 2.300 0.363 2.271

Panel B: Other benchmarks Baseline PT –

Adj. R2 selection 4.639 -0.352 4.485 2.149 0.192 2.140

All forecast functions 5.234 -1.340 5.060 3.277 -0.018 3.277

Equally weighted 4.485 -0.124 4.483 2.093 0.400 2.054 Inv. MSFE-

weighted 4.100 -0.146 4.097 1.924 0.358 1.890

Page 48: The Economic and Statistical Value of Forecast ... · 2For a special case, Aiolfi and Timmermann (2005) provide conditions under which the population MSFE of an equally-weighted

47

Table 8

Summary Statistics for Recursive Mean-Variance Portfolio Weights under a Variety of Forecasting Models/Combination Schemes

The table reports summary statistics for the weights solving the one-month forward mean-variance portfolio problem:

][21][max 11 ++ − tttt WVarWE tw

λ ,

where Wt+1 is end-of period wealth and λ is a coefficient of (absolute) risk aversion that trades-off mean and variance. The problem is solved recursively over the period 1985:01 – 2004:11 using in each month updated parameter estimates. The table shows means and standard deviations for recursive portfolio weights. Short-sales are ruled out. Results are shown for a few alternative values of the risk-aversion coefficient, λ.

Stocks Bonds Cash Mean Std. Dev. Mean Std. Dev. Mean Std. Dev.

MSIH(4,0)-VAR(1) 0.709 0.453 0.166 0.373 0.125 0.328 MSIAH(3,1) 0.401 0.486 0.259 0.434 0.340 0.469 VAR(1) 0.407 0.455 0.345 0.454 0.249 0.417 Neely-Weller benchmark 1.000 0.000 0.000 0.000 0.000 0.000 Pesaran-Timmermann ( 2R selection) 1.000 0.000 0.000 0.000 0.000 0.000 Forecast combinations Equally weighted 0.613 0.487 0.230 0.420 0.156 0.362

λ =

0.2

Inverse MSFE weights 0.593 0.491 0.240 0.427 0.167 0.373 MSIH(4,0)-VAR(1) 0.691 0.453 0.171 0.369 0.139 0.336 MSIAH(3,1) 0.383 0.475 0.255 0.426 0.362 0.469 VAR(1) 0.279 0.366 0.399 0.450 0.322 0.429 Neely-Weller benchmark 1.000 0.000 0.000 0.000 0.000 0.000 Pesaran-Timmermann ( 2R selection) 1.000 0.000 0.000 0.000 0.000 0.000 Forecast combinations Equally weighted 0.612 0.488 0.230 0.421 0.158 0.364

λ =

0.5

Inverse MSFE weights 0.589 0.491 0.241 0.425 0.169 0.374 MSIH(4,0)-VAR(1) 0.666 0.454 0.181 0.373 0.154 0.345 MSIAH(3,1) 0.353 0.456 0.260 0.423 0.387 0.470 VAR(1) 0.170 0.256 0.434 0.445 0.396 0.438 Neely-Weller benchmark 0.759 0.149 0.241 0.068 0.000 0.000 Pesaran-Timmermann ( 2R selection) 1.000 0.000 0.000 0.000 0.000 0.000 Forecast combinations Equally weighted 0.607 0.486 0.230 0.416 0.163 0.365

λ =

1

Inverse MSFE weights 0.581 0.489 0.246 0.424 0.173 0.376 MSIH(4,0)-VAR(1) 0.595 0.447 0.191 0.367 0.214 0.365 MSIAH(3,1) 0.288 0.405 0.271 0.414 0.441 0.464 VAR(1) 0.081 0.134 0.413 0.431 0.505 0.430 Neely-Weller benchmark 0.720 0.039 0.280 0.039 0.000 0.000 Pesaran-Timmermann ( 2R selection) 1.000 0.000 0.000 0.000 0.000 0.000 Forecast combinations Equally weighted 0.593 0.482 0.235 0.411 0.173 0.370

λ =

2

Inverse MSFE weights 0.563 0.490 0.255 0.425 0.182 0.380

Page 49: The Economic and Statistical Value of Forecast ... · 2For a special case, Aiolfi and Timmermann (2005) provide conditions under which the population MSFE of an equally-weighted

48

Table 9

Summary Statistics for Recursive Mean-Variance Portfolio Performances under a Variety of Forecasting Models/Combination Schemes

The table reports summary statistics for the 1-month portfolio return based on weights that solve the one-month mean-variance portfolio problem:

][21][max 11 ++ − tttt WVarWE tw

λ ,

where Wt+1 is end-of period wealth and λ is the coefficient of (absolute) risk aversion. The problem is solved recursively over the period 1985:01 – 2004:11 using in each month updated parameter estimates. Short-sales are ruled out. Results are shown for a few alternative values of the risk-aversion coefficient, λ. Mean-variance’ is the sample realized value of the mean-variance objective. Boldfaced values indicate the best performing model.

Mean Median 90% lower bound

90% lower bound

Sharpe ratio

Mean- variance

50% stocks, 50% bonds 1.010 1.015 0.954 1.050 0.153 1.009

1/N benchmark 1.008 1.011 0.970 1.035 0.163 1.007 MSIH(4,0)-VAR(1) 1.012 1.011 0.961 1.050 0.230 1.012

MSIAH(3,1) 1.008 1.005 0.961 1.046 0.128 1.008 VAR(1) 1.011 1.006 0.961 1.048 0.191 1.010

Neely-Weller benchmark 1.012 1.016 0.955 1.053 0.197 1.011

Pesaran-Timmermann ( 2R selection) 1.012 1.016 0.955 1.053 0.197 1.011

Forecast combinations Equally weighted 1.017 1.016 0.974 1.052 0.345 1.017

λ

= 0

.2

Inverse MSFE weights 1.017 1.016 0.977 1.052 0.353 1.017

MSIH(4,0)-VAR(1) 1.012 1.010 0.961 1.048 0.221 1.011 MSIAH(3,1) 1.008 1.004 0.963 1.046 0.137 1.008

VAR(1) 1.010 1.005 0.965 1.046 0.183 1.009 Neely-Weller benchmark 1.012 1.016 0.955 1.053 0.197 1.011

Pesaran-Timmermann ( 2R selection) 1.012 1.016 0.955 1.053 0.197 1.011

Forecast combinations Equally weighted 1.017 1.016 0.974 1.052 0.346 1.016

λ

= 0

.5

Inverse MSFE weights 1.017 1.016 0.977 1.052 0.354 1.016

MSIH(4,0)-VAR(1) 1.011 1.010 0.961 1.046 0.211 1.010 MSIAH(3,1) 1.008 1.004 0.963 1.046 0.138 1.007

VAR(1) 1.009 1.005 0.973 1.039 0.157 1.007 Neely-Weller benchmark 1.012 1.016 0.955 1.053 0.197 1.011

Pesaran-Timmermann ( 2R selection) 1.012 1.016 0.955 1.053 0.197 1.011

Forecast combinations Equally weighted 1.017 1.016 0.974 1.052 0.349 1.015

λ

= 1

Inverse MSFE weights 1.017 1.016 0.977 1.052 0.355 1.016

MSIH(4,0)-VAR(1) 1.010 1.008 0.963 1.044 0.199 1.007 MSIAH(3,1) 1.008 1.004 0.970 1.044 0.130 1.005

VAR(1) 1.007 1.005 0.976 1.030 0.118 1.005 Neely-Weller benchmark 1.011 1.016 0.955 1.052 0.181 1.007

Pesaran-Timmermann ( 2R selection) 1.012 1.016 0.955 1.053 0.197 1.011

Forecast combinations Equally weighted 1.017 1.016 0.977 1.052 0.356 1.014

λ

= 2

Inverse MSFE weights 1.017 1.015 0.977 1.052 0.359 1.014

50% stocks, 50% bonds 1.010 1.015 0.954 1.050 0.153 1.009

1/N benchmark 1.008 1.011 0.970 1.035 0.163 1.007

50% stocks, 50% bonds 1.010 1.015 0.954 1.050 0.153 1.008

1/N benchmark 1.008 1.011 0.970 1.035 0.163 1.007

50% stocks, 50% bonds 1.010 1.015 0.954 1.050 0.153 1.006

1/N benchmark 1.008 1.011 0.970 1.035 0.163 1.006

Page 50: The Economic and Statistical Value of Forecast ... · 2For a special case, Aiolfi and Timmermann (2005) provide conditions under which the population MSFE of an equally-weighted

49

Table 10

Sub-Sample Results for Recursive Mean-Variance Portfolio Performances The table reports summary statistics for the 1-month portfolio return based on weights that solve the one-month mean-variance portfolio problem. Results are broken down for four distinct periods: Jan. 1985 – Dec. 1989, Jan. 1990 – Dec. 1994, Jan. 1995 – Dec. 1999, and Jan. 2000 – Dec. 2004. Short-sales are ruled out. Results are shown for two alternative values of the risk-aversion coefficient, λ. Mean-variance’ is the sample realized value of the mean-variance objective. λ = 0.5 λ = 1 Mean Median Sharpe

ratio Mean-

variance Mean Median Sharpe ratio

Mean- variance

50% stocks, 50% bonds 1.012 1.015 0.172 1.010 1.012 1.015 0.172 1.009 1/N benchmark 1.010 1.012 0.196 1.009 1.010 1.012 0.196 1.008

MSIH(4,0)-VAR(1) 1.016 1.015 0.334 1.015 1.016 1.015 0.323 1.014 MSIAH(3,1) 1.011 1.007 0.178 1.010 1.011 1.007 0.181 1.009

VAR(1) 1.016 1.008 0.345 1.015 1.014 1.008 0.318 1.013 Neely-Weller benchmark 1.014 1.018 0.227 1.013 1.014 1.018 0.227 1.012

Pesaran-Timmermann ( 2R selection) 1.014 1.018 0.227 1.013 1.014 1.018 0.227 1.012

Forecast combinations Equally weighted 1.018 1.017 0.284 1.019 1.018 1.018 0.229 1.017

1985

- 1

989

Inverse MSFE weights 1.021 1.018 0.409 1.020 1.021 1.017 0.413 1.019 50% stocks, 50% bonds 1.008 1.012 0.141 1.012 1.008 1.012 0.141 1.007

1/N benchmark 1.007 1.009 0.152 1.006 1.007 1.009 0.152 1.006 MSIH(4,0)-VAR(1) 1.009 1.010 0.206 1.009 1.009 1.010 0.214 1.009

MSIAH(3,1) 1.009 1.004 0.204 1.009 1.009 1.004 0.210 1.008 VAR(1) 1.008 1.004 0.169 1.007 1.008 1.004 0.183 1.007

Neely-Weller benchmark 1.010 1.013 0.196 1.009 1.010 1.013 0.196 1.009

Pesaran-Timmermann ( 2R selection) 1.010 1.013 0.196 1.009 1.010 1.013 0.196 1.009

Forecast combinations Equally weighted 1.010 1.013 0.257 1.011 1.010 1.013 0.261 1.010

1990

– 1

994

Inverse MSFE weights 1.015 1.013 0.431 1.015 1.015 1.013 0.431 1.014 50% stocks, 50% bonds 1.019 1.028 0.378 1.018 1.019 1.028 0.378 1.017

1/N benchmark 1.014 1.020 0.390 1.014 1.014 1.020 0.390 1.013 MSIH(4,0)-VAR(1) 1.021 1.024 0.532 1.020 1.019 1.018 0.501 1.018

MSIAH(3,1) 1.010 1.004 0.203 1.010 1.010 1.004 0.200 1.009 VAR(1) 1.013 1.005 0.291 1.013 1.011 1.005 0.246 1.010

Neely-Weller benchmark 1.021 1.030 0.427 1.020 1.021 1.030 0.427 1.019

Pesaran-Timmermann ( 2R selection) 1.021 1.030 0.427 1.020 1.021 1.030 0.427 1.019

Forecast combinations Equally weighted 1.021 1.030 0.427 1.020 1.021 1.030 0.427 1.019

1995

- 1

999

Inverse MSFE weights 1.026 1.029 0.624 1.025 1.026 1.029 0.624 1.025 50% stocks, 50% bonds 1.001 1.001 -0.042 1.000 1.001 1.001 -0.042 0.999

1/N benchmark 1.001 1.007 -0.051 1.001 1.001 1.007 -0.051 1.000 MSIH(4,0)-VAR(1) 1.000 1.004 -0.076 0.999 1.000 1.004 -0.076 0.998

MSIAH(3,1) 1.003 1.002 -0.002 1.002 1.003 1.002 -0.001 1.001 VAR(1) 1.002 1.002 -0.018 1.001 1.001 1.003 -0.041 0.999

Neely-Weller benchmark 1.002 1.011 -0.020 1.001 1.002 1.011 -0.020 1.000

Pesaran-Timmermann ( 2R selection) 1.002 1.011 -0.020 1.001 1.002 1.011 -0.020 1.000

Forecast combinations Equally weighted 1.005 1.014 0.045 1.003 1.002 1.011 0.054 1.000

2000

- 2

004

Inverse MSFE weights 1.007 1.011 0.078 1.006 1.007 1.011 0.087 1.005

Page 51: The Economic and Statistical Value of Forecast ... · 2For a special case, Aiolfi and Timmermann (2005) provide conditions under which the population MSFE of an equally-weighted

50

Table 11

Recursive Mean-Variance Portfolio Performances under Transaction Costs The table reports summary statistics for the 1-month portfolio return based on weights that solve the one-month mean-variance portfolio problem. Short-sales are ruled out. Results are shown for two alternative values of the risk-aversion coefficient, λ. Mean-variance’ is the sample realized value of the mean-variance objective. Transaction costs are modeled as the function:

0|)ˆˆ| (max1

11

|ˆˆ|≠−

=−

−+−= ∑ j

tj

tj

wwf

n

j

jt

jtpt Iwwtc φφ

where p is a proportional transaction cost (set to 0.25 percent) and f is a fixed cost that has to be paid when transactions occur (f is set to 0.01 percent).

Mean Median 90% lower bound

90% lower bound

Sharpe ratio

Mean- variance

MSIH(4,0)-VAR(1) 1.010 1.009 0.958 1.047 0.176 1.009 MSIAH(3,1) 1.006 1.003 0.963 1.043 0.069 1.005

VAR(1) 1.007 1.004 0.961 1.042 0.104 1.006 Neely-Weller benchmark

1.012 1.016 0.955 1.053 0.197 1.010

Pesaran-Timmermann ( 2R selection)

1.012 1.016 0.955 1.053 0.197 1.011

Forecast combinations Equally weighted 1.014 1.017 0.952 1.055 0.236 1.012

λ

= 0

.5

Inverse MSFE weights 1.013 1.013 0.972 1.050 0.280 1.014

MSIH(4,0)-VAR(1) 1.009 1.008 0.958 1.045 0.162 1.008 MSIAH(3,1) 1.006 1.003 0.961 1.043 0.069 1.004

VAR(1) 1.006 1.004 0.971 1.036 0.086 1.005 Neely-Weller benchmark

1.012 1.016 0.955 1.053 0.197 1.010

Pesaran-Timmermann ( 2R selection)

1.012 1.016 0.955 1.053 0.197 1.010

Forecast combinations

Equally weighted 1.013 1.016 0.954 1.053 0.241 1.011

λ

= 1

Inverse MSFE weights 1.014 1.013 0.973 1.051 0.281 1.013

50% stocks, 50% bonds 1.010 1.015 0.954 1.050 0.153 1.009

1/N benchmark 1.008 1.011 0.970 1.035 0.163 1.007

50% stocks, 50% bonds 1.010 1.015 0.954 1.050 0.153 1.008

1/N benchmark 1.008 1.011 0.970 1.035 0.163 1.007